Sunteți pe pagina 1din 130

Accepted Manuscript

Thermomechanical processing of advanced high strength steels

Jingwei Zhao, Zhengyi Jiang

PII: S0079-6425(18)30012-4
DOI: https://doi.org/10.1016/j.pmatsci.2018.01.006
Reference: JPMS 492

To appear in: Progress in Materials Science

Received Date: 8 June 2017


Revised Date: 18 January 2018
Accepted Date: 18 January 2018

Please cite this article as: Zhao, J., Jiang, Z., Thermomechanical processing of advanced high strength steels,
Progress in Materials Science (2018), doi: https://doi.org/10.1016/j.pmatsci.2018.01.006

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 proof before it is published in its final 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.
Thermomechanical processing of advanced high strength steels
Jingwei Zhao*, Zhengyi Jiang*

School of Mechanical, Materials, Mechatronic and Biomedical Engineering, University of


Wollongong, NSW 2522, Australia

Abstract

Advanced high strength steels (AHSSs) are regarded as the most promising materials for
vehicles in the 21st century. AHSSs are complex and sophisticated materials, with
microstructures being controlled by precise thermomechanical processing (TMP) technologies.
TMP is an established and strategic method for improving the mechanical properties of AHSSs
through control of microstructures and is among the most important industrial technologies for
producing high quality AHSSs with the necessary mechanical properties. This article aims to
provide a comprehensive review of recent progress in TMP of AHSSs, with focus on the
processing-microstructure-property relationships of the processed AHSSs. We first present an
introduction to the background of the TMP of AHSSs. Then, the recent progress and the latest
achievements in TMP of the first, second and third generations of AHSSs and Nano Hiten steels
are reviewed in detail, and the mechanisms of the TMP-induced microstructural evolution and
mechanical properties variation are addressed and discussed. The present review concludes with
a summary on the TMP of AHSSs currently under development, and also offers an outlook of
the future opportunities which will inspire more in-depth research and eventually advance
practical applications of this innovative field.

Keywords: Advanced high strength steels; Thermomechanical processing; Microstructure;


Mechanical properties; Automotive applications

* Corresponding authors.

E-mail addresses: jwzhaocn@gmail.com (J. Zhao), jiang@uow.edu.au (Z. Jiang)

1
Contents
1. Introduction .......................................................................................................................... 4

2. Thermomechanical processing of the first generation AHSSs ................................................ 8

2.1 Thermomechancal processing of DP steels ...................................................................... 8

2.2 Thermomechancal processing of TRIP steels ................................................................. 20

2.3 Thermomechancal processing of CP steels..................................................................... 33

2.4 Thermomechancal processing of MART steels .............................................................. 37

3. Thermomechanical processing of the second generation AHSSs .......................................... 43

3.1 Thermomechancal processing of TWIP steels ................................................................ 43

3.2 Thermomechancal processing of AUST.SSs .................................................................. 54

3.3 Thermomechancal processing of L-IP steels .................................................................. 68

4. Thermomechanical processing of the third generation AHSSs ............................................. 75

4.1 Quenching and partitioning ........................................................................................... 79

4.2 Martensite-to-austenite reversion treatment.................................................................... 83

4.3 Quenching-partitioning-tempering ................................................................................. 89

4.4 Dual stabilisation heat treatment .................................................................................... 90

5. Thermomechancal processing of Nano Hiten steels ............................................................. 91

6. Summary and outlook ....................................................................................................... 100

Acknowledgement ................................................................................................................ 102

References ............................................................................................................................ 103

2
Nomenclature

α temperature-dependent parameter
Ac1 austenitisation starting temperature during heating
Ac3 austenitisation finishing temperature during heating
Ae3 equilibrium austenite-to-ferrite transformation starting temperature
Ar1 austenite-to-ferrite finishing temperature during cooling
Ar3 austenite-to-ferrite starting temperature during cooling
β temperature-dependent parameter
b Burgers vector
df average ferrite grain size
dp average diameter of precipitates
 g grain refinement strengthening
 p precipitation strengthening
ε plastic strain
f  volume fraction of martensite
fp volume fraction of precipitates
K constant
ky constant
Mf finishing temperature of martensitic transformation
Ms starting temperature of martensitic transformation
n strain hardening exponent
n' a fixed exponent
AHSS advanced high strength steel
AUST.SS austenitic stainless steel
BCC body-centred cubic
CP complex phase
DIFT deformation-induced ferrite transformation
DP dual phase
DRV dynamic recovery
DRX dynamic recrystallisation
DSHT dual stabilisation heat treatment
FCC face-centred cubic
HSLA high-strength low-alloy
L-IP lightweight steel with induced plasticity
MART martensitic
Q&P quenching and partitioning
Q-P-T quenching-partitioning-tempering
TMP thermomechanical processing
TRIP transformation-induced plasticity
TWIP twinning-induced plasticity
UTS ultimate tensile strength
YS yield strength

3
1. Introduction
Reducing the weight of a vehicle is one most direct and effective way to improve fuel
efficiency and reduce greenhouse gas emissions. However, there are potential safety problems
that need to be considered. For example, an average of 4.9% fuel economy improvement could
be achieved for every 10% reduction in total vehicle weight [1] but a 3%-4.5% increase in the
safety risk could be caused with a decrease of 100 kg in the weight of a car [2,3]. Automotive
manufacturers must balance the objectives of both improved fuel economy and crash safety. In
an attempt to respond such objectives, the steel industry has been developing new advanced
high strength steel (AHSS) with unique metallurgical properties and processing methods of
which enable the automotive industry to meet requirements for safety, efficiency, emissions,
manufacturability, durability and quality at relatively low cost [4,5]. Many groups all over the
world are conducting research on these new steel grades to understand better their applications
and to continue tailoring unique sets of characteristics. It is possible to save 25% of the weight
and 14% of the cost when conventional steels in a four-door car body are replaced by AHSSs
[6]. It has been reported that the energy consumption in the transportation sector will decline
from 26.7 quadrillion Btu in 2012 to 25.5 quadrillion Btu in 2040 due to a significant decline in
energy consumption by light-duty vehicles [7], which will decrease the emission of greenhouse
gas and benefit global climate change. Based on an environmental case study reported by the
World Steel Association [8], the use of every 1 kg of AHSSs in a five-passenger family car
could achieve a total life cycle saving of 8 kg greenhouse gas (reported as CO2 equivalents),
which corresponds to a 5.7% reduction in greenhouse gas emissions over the full life cycle of
the vehicle. It is estimated that the ratio of AHSSs to the total automotive steels would be
increased from 7% in 2009 to 28%-36% in 2020 [9]. AHSSs have a bright future, and have been
regarded as the most promising materials for vehicles in the 21st century due to the unique
combination of excellent performance and competitive cost.
Based on the different development stages, AHSSs are generally divided into three
generations [10]. Fig. 1 shows the locations of three generations of AHSS grades in an
elongation versus tensile strength chart [11]. The first generation AHSSs include dual phase
(DP) steel, transformation-induced plasticity (TRIP) steel, complex phase (CP) steel and
martensitic (MART) steel. These steel grades have higher strength and better ductility without
significant increase in cost as compared to the conventional high strength steels, and have been
well developed and used in many applications throughout the automotive industry. The second
generation AHSS grades which include twinning-induced plasticity (TWIP) steel, austenitic
stainless steel (AUST.SS) and lightweight steel with induced plasticity (L-IP) are located in the
top right corner of the elongation-strength chart, indicating these steels possess superior

4
mechanical properties. These steels are extremely strong and formable, and exhibit an excellent
combination of strength and formability to ensure they can be used to provide extraordinary
mass reduction for difficult-to-form parts, but these AHSS grades are highly alloyed with high
cost alloying elements such as chromium and nickel, resulting in significant production cost
increase, and their widespread adoption is hampered. The third generation AHSSs have a better
combination of strength and ductility than the first generation one, and can be produced with a
lower cost than that of the second generation one. The third generation grades are currently
under research and development. They are regarded as future opportunities for automotive and
other structural applications due to their improved strength and ductility relative to the first
generation and more affordable than the second generation.

Second generation AHSS

Third generation AHSS

First generation AHSS

Fig. 1. Locations of three generations of AHSS grades in an elongation versus tensile


strength chart [11].

In addition to the above mentioned three generatioins of AHSSs, there is one more group
which is usually referred to as Nano Hiten steels. The term “Nano” is abbreviated from the
phrase of “New Application of Nano Obstacles for Dislocation Movement”. Nano Hiten steels
were developed by the JFE Steel Corporation, Japan, in the early 2000s by precipitating nano-
sized carbides in Ti-Mo bearing high-strength low-alloy (HSLA) steels with the purpose of
improving the steels’ hole expansionability [12,13]. For steels under the same or similar tensile
strength grade, Nano Hiten steel can exhibit a higher hole expansion ratio and elongation as
compared to conventional steels, as shown in Fig. 2 [14]. Nano Hiten steels have been used in
automotive suspensions and crashworthy equipment, and their applications are expected to
increase in the future due to the attractive combination of the high strength and high hole
expansionability.

5
Fig. 2. Comparison of the depences of hole expansion ratio on elongation in 780 MPa
grade hot rolled steels [14].

AHSSs for auto-making need to have excellent mechanical properties so that they can be
used properly in the automotive body structure [15]. For a given AHSS with specified chemical
compositions, the mechanical properties are primarily a function of the steel’s microstructure
which is heavily dependent on the processing conditions such as temperature, strain, strain rate,
deformation mode and cooling method [16]. Thermomechanical processing (TMP), a
combination of deformation and heat treatment, is an established and strategic method for
improving the mechanical properties of AHSSs through control of their microstructure. TMP is
among the most important industrial technologies for producing high quality AHSSs with the
required mechanical properties. It improves the microstructural features in order to realise the
AHSS products fitting the requirements imposed by modern automotive technology. TMP is a
metallurgical process that integrates work hardening and heat treatment into a single process
[17]. It is principally applied to the processing of metallic alloys, but it has also found
applications in the manufacturing of ceramics, polymers and many of their combinations [16].
The TMP method was first introduced into commercial production of C-Mn steel plates in the
1950s by means of controlled rolling [18], and increasingly it has become an important
technique in the process design of controlled rolling, cooling and direct quenching of many steel
products, such as plates, sheets, strips, beams, bars, wires, pipes and rails [19-23].
The key feature of TMP is its sophisticated combination of well defined deformation
operations and well defined heat treatment in a single stage to control the microstructure of
materials. Fig. 3 shows a schematic illustration in which the conventional rolling including the
normalising treatment of steel plates is compared with the TMP method [24]. It can be seen that
no offline heat treatment is required in the TMP as compared with the conventional

6
manufacturing process. In the hot rolling of conventional high strength steels, such as HSLA
steels, TMP implies various types of controlled rolling where temperature, strain rate and
reductions are carefully controlled to produce a fully recrystallised or fully pancaked austenite
microstructure at the end of finish rolling [25]. The combination of microalloying and suitable
TMP provides an efficient approach to control austenite recrystallisation, and as a result, fine
ferrite grains can be obtained by the refined mother phase of austenite during controlled rolling
in combination with the following accelerated cooling. In the manufacturing of steels, TMP
saves energy by minimising or even eliminating the heat treatment process after hot working,
thus increasing the productivity for high grade steels. TMP has been thought to be central to the
concept of controlled processing that has transformed steelmaking into a modern, cost-effective,
and high-quality manufacturing industry [24,26].

Normalising TMP
Temperature

Austenite

Accelerated cooling

Austenite→Ferrite

Time
Fig. 3. Comparison of the conventional normalising treatment with TMP [24].

Conventional low- to high-strength steels, including mild, interstitial-free, bake-hardenable


and HSLA steels, can be processed by the existing mature TMP technologies because they have
relatively simple microstructures [27,28]. The term “AHSSs”, instead, refers to a new group of
high strength steels which include three generations of AHSSs and Nano Hiten steels. A lot of
AHSSs are simultaneously alloyed with significant amounts of ferrite and austenite stabilising
elements in combination with microalloying and other transformation retarding elements [29].
Such complicated alloying leads to difficulties in the control of microstructural evolution in the
processing of AHSSs because the final microstructure of AHSSs is usually a composite with
different combinations of ferrite, martensite, bainite, austenite and precipitates. As a result, it
may not be suitable to simply apply the existing TMP routes to AHSSs. With the ongoing
growth of AHSS production, the quality requirements for AHSS products are becoming
increasingly strict, and will continue to toughen [30]. Under this circumstance, the contribution
of each manufacturing stage to the overall product quality must be carefully tailored, and
different processing strategies should be adopted for specified AHSS grades. As an important

7
stage prior to the industrial rolling production, research into the TMP of AHSSs in the
laboratory environment becomes essential and significant for the optimal design of processing
conditions by precise control of temperature, strain, strain rate, heat treatment and cooling
parameters. Each type of AHSS has a unique application in vehicles, and specified TMP
technologies should be developed to produce high quality AHSS products where they might be
the best employed to meet mechanical property demands for the automotive parts.
Although a large amount of research has been done into various aspects of AHSSs such as
chemical composition design [31], microstructural features [32], processing [33], mechanical
properties [34], formability [35], heat treatment [36], hydrogen embrittlement [37-39] and
crashworthiness [40], there is still a lack of research into the progress in TMP of AHSSs. At
present, only limited AHSSs for the automotive industry are being produced in the production
lines of giant steel suppliers like Baosteel (China), JFE Steel Corporation (Japan) and POSCO
(South Korea) due to the difficulties in the processing of these newly developed steels. Research
into TMP of AHSSs is still in progress with the growing demands for AHSSs in automotive and
other manufacturing industries, and many problems have arisen in the processing of these steels
which still need to be solved with the increasing requirements for the excellent mechanical
properties and reduced cost of the processed AHSS products. The purpose of this review is to
assess the progress in TMP of AHSSs, while focusing on the processing-microstructure-
property relationships of the processed AHSSs. After the introduction in Section 1, the TMP of
the first, second and third generations of AHSSs will be systematically reviewed in Sections 2, 3
and 4, respectively. In Section 5, the progress in TMP of Nano Hiten steels will be addressed
and discussed. In these sections, the latest achievements in TMP of AHSSs will be introduced in
detail, and the mechanisms of TMP-induced microstructural evolution and mechanical
properties variation will be highlighted and discussed. Finally, authoritative comments will be
provided to summarise this review, and the future outlook for development of the TMP of
AHSSs will be presented. This review will provide a comprehensive literature for those who are
developing, making, using and designing AHSSs and their TMP technologies, but due to the
fast pace of research in this area, it is impossible to examine all the related studies being
conducted in the world, and some newly published articles might inevitably be missed.

2. Thermomechanical processing of the first generation AHSSs


2.1 Thermomechancal processing of DP steels
DP steels consist of a soft ferrite matrix and hard martensite islands [41]. A schematic of the
microstructure of DP steels is shown in Fig. 4 [42]. Fig. 5 shows an actual optical microscope
microstructure of DP steel, in which the white phase is martensite and the rest is ferrite [43].

8
The soft ferrite generally distributes continuously for many DP steel grades up to DP 780, but
the ferrite may become discontinuous as the volume fraction of martensite exceeds 50% [11].
Ferrite provides DP steels with excellent ductility, whereas martensite islands contribute to
strength. The higher the volume fraction of the martensitic phase, the higher the strength of the
steels will be. Such microstructural characteristics enable DP steels to achieve an ultimate
tensile strength (UTS) in the range of 500-1200 MPa [4].

Ferrite

Martensite

Fig. 4. Schematic microstructure of DP steels [42].

Fig. 5. Optical microscope microstructure of a DP steel [43].

DP steels use a low C-Mn-Si system as the basic chemical composition. Alloying elements
determine the phase constituents and microstructural characteristics of DP steels, and therefore
their final mechanical properties after processing. Carbon in DP steels acts as an austenite
stabiliser, strengthens martensite, determines the phase distribution, and affects greatly the
variation of mechanical properties. The carbon content in DP steels is generally used in the
range of 0.06-0.15 wt.%. Manganese, used between 1.5-2.5 wt.%, also stabilises austenite, and
is a ferrite solid solution strengthener and a ferrite formation retarder. Silicon in DP steels plays
a role of promoting ferritic transformation. The addition of other alloying elements such as
chromium, molybdenum, vanadium and niobium individually or in combination helps to
stabilise austenite, modify microstructure and improve mechanical properties. Chromium and
molybdenum usually do not exceed 0.4 wt.%, and can retard pearlite and bainite formation.
9
Vanadium is a precipitation strengthener and can effectively refine microstructure. Generally,
the content of vanadium in DP steels should not surpass 0.06 wt.%. Niobium, used up to 0.04
wt.%, is also a precipitation strengthener. It can reduce martensite transformation temperature,
refine microstructure and promote ferrite transformation from non-recrystallised austenite [4,44].
The addition of these alloying elements should be carefully balanced, not only to produce
unique mechanical properties, but also to maintain the generally good resistance spot welding
capability.
DP steels exhibit continuous yielding, i.e. there are no apparent yield points in the stress-
strain curves after tensile tests. Continuous yielding is caused by the presence of internal
stresses which allow dislocations to remain mobile, i.e. dislocations are not pinned by interstitial
atoms during deformation [6,45]. The soft ferrite phase in DP steels is exceptionally ductile and
absorbs strain around the martensitic islands, enabling a unique excellent elongation and higher
UTS than that of conventional steels with similar yield strength (YS). Fig. 6 compares the
engineering stress-strain curves for DP steel to a HSLA steel with similar YS of 350 MPa [42].
It is clear that DP steel exhibits a higher initial work-hardening rate, higher UTS and lower
YS/UTS ratio than that of the HSLA steel with similar YS. Recently Tasan et al. [46] published
a review paper on the microstructure-oriented processing and micromechanically-guided design
of DP steels, and Perzyński et al. [47] reviewed the modelling of the fracture behaviour of DP
steels.

Fig. 6. Engineering stress-strain curves of DP and HSLA steels with a similar YS [42].

During the processing of ferrite plus martensite DP steels, austenite is usually first suffered
controlled-cooling to generate a portion of ferrite before an accelerated cooling to transform the
remaining austenite to martensite. Due to such a processing characteristic, small amounts of
other phases including bainite and/or retained austenite may exist in the final DP steel products
[42]. In practice, DP steels are produced by controlled cooling from the austenite region for hot-
rolled products, or from the “ferrite + austenite” region during an intercritical annealing

10
treatment for continuously annealed cold-rolled and hot-dip coated products to transform some
austenite to ferrite before accelerated cooling transforms the remaining austenite to martensite.
Fig. 7 shows a schematic diagram of the processing of DP steels through controlled cooling
from austenite and “ferrite + austenite” regions, in which Ac1 and Ac3 are respectively the
austenitisation starting and finishing temperatures during heating, and Ms and Mf are the starting
and finishing temperatures of martensitic transformation, respectively.

Ac3 A
F

Ac1

Ferrite
Pearlite
Bainite
Ms
M
Mf F

Fig. 7. Schematic diagram of the processing of DP steels. A: austenite,


F: ferrite, M: martensite.

Generally, the tensile strength of DP steels is linearly proportional to the amount of


martensite in the structure, i.e. the higher the martensite volume fraction, the higher the tensile
strength will be [48-50], but such a trend is not always valid, and a critical value of martensite
volume fraction exists for achieving the highest tensile strength of DP steels. Pouranvari [51]
found that the tensile strength of DP steel first increased as the martensite volume fraction is
increased, and then reached the maximum tensile strength at the martensite volume fraction of
50%. The work of Sudersanan et al. [52] showed that the critical volume fraction of martensite
for achieving the highest tensile strength was 63% in DP steel. An increase in the volume
fraction of martensite within an appropriate range can lead to an enhancement of strength of DP
steels on the basis of the ferrite-martensite microstructure only, but the ductility of DP steels
will be sacrificed by such simple increase of the volume fraction of martensite [53]. Another
way to increase the strength of DP steels is to refine the microstructure. It has been reported that
grain refinement of DP steels offers a promising way to strengthen as the increase in strength is
not accompanied by a loss in strain hardenability or total elongation [54-57]. DP steels with
refined ferrite grains and fine uniform dispersion of martensite in the ferrite matrix provide the

11
best combination of strength and ductility compared with those have blocky ferrite-martensite
and martensite islands along the grain boundaries of polygonal ferrite grains [58,59]. The grain
refinement of DP steels can be achieved by means of equal channel angular pressing [56,60] or
accumulative roll bonding [61] but both methods have limitations in the sample size and shape,
therefore restricting their applications in large scale productions. Alternatively, TMP provides
an effective way for grain refinement of DP steels with cost effectiveness and ease of
manufacturing in industry.
Karmakar et al. [62] conducted a comparative study on the development of ultrafine-grained
ferrite/martensite DP structures in a low-carbon microallyed steel using three TMP schedules of
(i) “cooling + intercritical deformation”, (ii) “heating + intercritical deformation”, and (iii)
“warm-deformation + intercritical annealing”, as shown in Fig. 8a, b and c, respectively. In the
three schedules, the temperatures were determined on the basis of the equilibrium austenite-to-
ferrite transformation starting temperature (Ae3=1133 K, i.e. 860 ℃), and the austenite-to-ferrite
starting (Ar3=1087 K, i.e. 814 ℃) and finishing (Ar1=918 K, i.e. 645 ℃) temperatures during
cooling. The soaking temperature of 1273 K was selected because at this temperature uniform
austenite structure with a unimodal distribution of austenite grain sizes could be obtained after 5
min soaking [63]. Elements of niobium and titanium present in precipitate form at this
temperature, thus restricting the growth of austenite during soaking [64]. It has been proven that
fine austenite grain size with the absence of niobium dissolved in austenite benefits the
deformation-induced ferrite transformation (DIFT), which is also known as dynamic strain-
induced transformation of austenite-to-ferrite, in the following deformation process [65-67]. It
should be noted however that the austenitising temperature varies, and depends on the chemical
compositions and the targeted final microstructural constituents. Santos et al. [68] found that a
more homogeneous ferrite microstructure after hot rolling can be obtained after austenitisation
at 900 ℃ than that austenitised at 1200 ℃. The work of Salehi et al. [69] on a low alloy steel
showed that austenisation at temperatures higher than 1000 ℃ may cause an increase in the
volume fraction of bainite after hot rolling, and that soaking at 900 ℃ for 30 min could acquire
the optimised DP structure which consists of a fine polygonal ferrite matrix and about 15% fine
martensite particles. This means that soaking parameters should be determined on the basis of
the studied steels, the targeted microstructures and the processing conditions.

12
Fig. 8. Different TMP schedules used for producing ultrafine-grained DP structures: (a) intercritical
deformation after cooling down the samples from the soaking temperature to the deformation
temperatures, (b) intercritical deformation after heating up the samples from an intermediate
temperature to the deformation temperatures, and (c) warm-deformation followed by intercritical
annealing [62]. WQ: water quenching.

Deformation temperature has a significant effect on microstructural refinement after hot


deformation for the schedule (i), as shown in Fig. 8a. Above Ar3, martensite will dominate the
final microstructure [63], and DIFT is responsible for the grain refinement by dynamic
recrystallisation (DRX) of ultrafine ferrite grains [70,71]. In the intercritical temperature range
between Ar3 and Ar1, the microstructural evolution mechanism changes from ferrite recovery and
continuous DRX to deformed and recovered structure with a high fraction of low-angle
boundaries. Intercritical deformation of the schedule (i) can provide optimal grain refinement
(1.4-2.7 μm grain size), the highest fraction of high-angle boundaries (75%-84%), and
beneficial gamma-fibre texture component of <111>//normal direction by virtue DIFT and
continuous DRX [63]. A slight coarser ferrite grain (coarser by 0.5-1.0 μm) and higher amounts
of martensite will be caused during the schedule (ii) (Fig. 8b) as compared to that processed by
the schedule (i) due to the strain partitioning in ferrite and adiabatic heating during the
deformation of the samples [72]. This schedule is not suitable because the generated high
fraction of martensite exceeds 30% which is not desirable even though a uniform distribution of

13
martensitic islands can be obtained [62]. As compared with the schedules (i) and (ii), DIFT and
DRX are not essential for the formation of ultrafine-grained structure for the schedule (iii) (Fig.
8c). The warm deformation followed by intercritical annealing can also develop similar
structures that are composed of ultrafine-grained ferrite (1.4-2.7 μm) with an optimal martensite
fraction of 20%-30% and higher than 70% high-angle boundaries through the static
recrystallisation mechanism [62]. The schedule of warm deformation plus intercritical annealing
is easier to control, and the effect of adiabatic heating is the least in this schedule. However, this
schedule involves low-temperature deformation and high heating and cooling rates in the
intercritical annealing, which challenges the capability of the hot rolling mill and requires a
precise control over the annealing treatment. As a result, application difficulties of the warm
deformation schedule for industrial manufacturing line will be increased.
Mazaheri et al. [73] proposed a cold rolling followed by intercritical annealing method for
fabricating ultrafine-grained ferrite-martensite microstructure with simultaneously improved
mechanical properties. In this method, steel was first austenitised at 880 ℃ for 60 min, and then
intercritically annealed at 770 ℃ for 100 min followed by water quenching. The water-
quenched steel was cold rolled to obtain a duplex ferrite-martensite structure after a reduction of
80% followed by a further annealing plus water quenching to generate ultrafine/nanoferrite-
carbide aggregate prior to the final intercritical annealing, as illustrated in Fig. 9a (DPI). For
comparison, they also employed fewer steps after cold rolling by eliminating the annealing and
water quenching operations, as shown in Fig. 9b (DPII). In order to investigate the effects of
intercritical annealing time and temperature on the microstructural evolution and mechanical
properties, the cold-rolled steel was also annealed at 770 ℃ for 10 min (DPIII) and 790 ℃ for 8
min (DPIV). In the TMP, intercritical annealing at 770 ℃ for 100 min followed by water
quenching is used to generate a duplex microstructure composing of a ferrite matrix and
martensite islands. The high reduction cold rolling induces a non-uniform strain distribution
between softer ferrite and harder martensite, i.e. higher strain is introduced in the softer ferrite
matrix due to the existence of the harder martensite phase, which plays a key role in the
formation of fine ferrite-martensite microstructure in the subsequent intercritical annealing
treatment. In comparison with the schedule DP II, DPI will cause coarser ferrite grain size and
higher volume fraction of martensite due to the additional annealing at 600 ℃ for 20 min.
Austenite preferentially nucleates at the carbide particles located at ferrite grain boundaries. A
combination of large population of ferrite grain boundaries and carbide particles will increase
the austenite nucleation site density and austenite volume fraction, which contribute to a higher
volume fraction of martensite after water quenching [74,75]. The larger ferrite grain size of DPI
is attributed to the lower stored energy in the initial microstructure of the DP I prior to the final

14
intercritical annealing [73]. Therefore, the schedule DPI should be excluded due to the
complicated processing steps and undesirable final microstructure obtained. The holding time
and temperature during intercritical annealing treatment have a significant effect on final
microstructural characteristics. The volume fraction of martensite increases with increasing the
intercritical annealing holding time and temperature due to the increased amount of austenite
which will transform to martensite upon quenching in water [76-79]. The final ferrite grain size
increases and decreases slightly with increasing the intercritical annealing holding time and
temperature, respectively. Such variation of ferrite grain size is thought to be caused by the
volume fraction of martensite, i.e. the higher amount of martensite phase, the finer the ferrite
grains will be caused [80]. Finally, the ultrafine-grained DP steel with an average grain size of
2.13 μm was obtained by the schedule DPIV with a good combination of ultrahigh tensile
strength of about 1430 MPa and adequate total elongation of about 12.6%. The TMP schedule
of DPIV is effective for grain refinement of DP steels, but the improved tensile strength and
elongation are still limited due to the limitation of grain size refinement.

(a) (b)

Fig. 9. TMP schedules performed on a low-carbon steel to develop ultrafine-grained


ferrite-martensite DP structure [73]. CR: cold rolling, WQ: water quenching.

To further refine the microstructure and then improve mechanical properties, Mazaheri et al.
[81] modified the TMP schedule of DPIV (Fig. 9b) by increasing the intercritical annealing
temperature after cold rolling. They found that the ferrite grain size showed a decreasing trend
with the increase of intercritical annealing temperature. The refined ferrite grains are thought to
be caused by the increased volume fraction of martensite phase [80]. An increase in the
intercritical annealing temperature enhances both the YS and UTS, but induces the decreased
elongation due to the increased amount of martensite in the final microstructure [81-83], as
shown in Fig. 10. They concluded that the ultrafine-grained ferrite-martensite DP
microstructure with average grain size of 1.42 μm could be obtained at the intercritical

15
annealing temperature of 830 ℃ for 8 min followed by water quenching, and the tensile strength
could achieve 1600 MPa, and the total elongation could be slightly enhanced from 12.6% to
about 13% with a uniform elongation of about 7%.

(a) (b)

Fig. 10. Dependences of (a) YS and UTS and (b) uniform and total elongation on the
intercritical annealing temperature [81].

Mukherjee et al. [84] employed a TMP route considering the mechanism of DIFT to refine
the ferrite grain size of four low-carbon steels using Gleeble and hot torsion simulations, as
shown in Fig. 11. DIFT is the most efficient method of grain refinement for plain low-carbon
steel without addition of microalloying elements. This is because that the grain growth rate of
plain low-carbon steel is not controlled by the TMP during the rolling process. To obtain DIFT,
rapid cooling (e.g. 40 ℃/s and above) is essential to generate a strongly undercooled metastable
austenite microstructure, and the following hot deformation is generally performed in a
temperature range that is 25-100 ℃ above Ar3 with a heavy reduction [84]. Significant
undercooling under Ae3 provides a sufficiently high driving force for ferrite nucleation, which
promotes the generation of ultrafine ferrite grains. Using the DIFT schedule of Fig. 11,
Mukherjee et al. [84] successfully produced ultrafine ferrite grains with a size of 1-2 μm for all
investigated steels when a true strain of 0.6 was applied to an austenitic microstructure with
austenite grain size in the range of 10-20 μm.

16
Fig. 11. Schematic illustration of TMP schedule with DIFT [84].

In DIFT, austenitisation obtains a uniformly distributed prior austenite structure with an


optimal austenite grain size. Beladi et al. [85] indicated that the distribution of ferrite grains
after DIFT was more homogenous for the fine prior austenite grain size compared with the
coarse austenite. In contrast, Hurley et al. [86] argued that the coarser prior austenite grains
suppressed the formation of grain-boundary proeutectoid ferrite and thus enhanced DIFT.
Mukherjee et al. [84] thought a fine prior austenite grain size (about 10 μm) was required to
produce a predominantly ultrafine ferrite microstructure of the niobium and molybdenum-
niobium microalloyed steels, while a predominantly ultrafine ferrite microstructure could also
be obtained from the coarser prior austenite grain size of about 27 μm for the molybdenum-
alloyed steel. Therefore, there is still no consistent agreement upon the role of prior austenite
grain size in DIFT for different DP steels and TMP schedules. A critical strain, under which no
DIFT will occur, is required for ultrafine ferrite formation [87]. Hong et al. [88] found that the
critical strain for the formation of ultrafine ferrite decreased with a decrease in the deformation
temperature. This finding was confirmed by Beladi et al. [89] that the critical strain increased
with an increase of the prior austenite grain size and deformation temperature, and decreased
with an increase of the post-deformation cooling. At strains higher than the critical strain value,
the final microstructure did not change significantly. The strain rate could play an important role
in DIFT and microstructure. The work of Beladi et al. [89] showed that an increase in strain rate
induced finer ferrite grain size after DIFT. Hong and Lee [90] found that employing a higher
strain rate enhanced DIFT. Yang and Wang [91] reported that the strain rate played a very
important role by controlling the holding time and activation energy required by DIFT when the
steel was processed under the same strain level. They concluded that the strain rate affected the
amount of deformation required to achieve DIFT. Rapid cooling after DIFT deformation is the
key step in obtaining martensite as a secondary transformation product and attaining the desired
fine-grained DP microstructure [84,91].

17
In addition to the aforementioned methods, a variety of other TMP routes were performed to
refine the microstructure and achieve grain sizes of a few microns or even approximately 1 μm
[92-94]. The researchers pursued grain refinement of DP steels to realise improved mechanical
properties, especially strength and toughness, without costly alloying additions. However, the
uniform elongation achieved in ultrafine-grained DP steel is smaller than that of the coarse-
grained structure due to weak dislocation activity within the ultrafine grains [54,56,95,96]. In
order to improve the uniform elongation of ultrafine-grained DP steels, a number of different
TMP routes are being pursued to develop the uniform dispersion of fine-carbide particles and/or
bimodal ferrite grain structure. The bimodal ferrite grain size distribution observed after hot
deformation is attributed to the heterogeneous distribution of composition, precipitation and
strain, which induce local difference in austenite recrystallisation, grain growth and austenite-to-
ferrite transformation kinetics [97-99]. Karmakar et al. [100] developed a bimodal ultrafine-
grained DP steel with a tensile strength of 744 MPa and a uniform elongation of 9.3% by rapid
intercritical annealing following warm rolling of a ferrite-pearlite steel. They found that fine
austenite grains developed during rapid annealing and transformed into fine-ferrite grains after
cooling. Coarse-ferrite grains resulted from the recrystallisation and growth of deformed ferrite.
Bimodal grain structures showed better tensile ductility, especially in terms of uniform
elongation, along with satisfactory strength in comparison to the ultrafine-grained steel. Azizi-
Alizamini et al. [101] reported that bimodal grain size distribution of ultrafine-grained ferrite
and carbide particles could be obtained in low-carbon steel through full water quenching,
followed by intercritical quenching, then a 50% reduction in cold rolling and long-term
annealing at a low temperature (annealing at 525 ℃ for 1200 min). The main mechanisms by
which the heterogeneous microstructure could be created were the concurrent recrystallisation
of ferrite and martensite combined with carbide precipitation in the martensitic regions.
Employing this method, a uniform elongation of 14% was achieved but the tensile strength of
550 MPa was not as high. In addition, the very long annealing time (1200 min) challenges the
practical applications of this TMP route in industry.
In order to achieve high uniform elongation and high tensile strength, Wang et al. [102]
proposed a TMP route to tailor the microstructure to duplex-sized recrystallised ferrite grains
composed of submicron and several micron grains coupled with dispersive nanoscale carbides
using appropriate annealing of cold-rolled ferrite-martensite DP structure of a low-carbon steel.
They obtained a high uniform elongation of 17.8% with a tensile strength of 788 MPa.
Obviously, the tensile strength is still not high enough to ensure the treated steel can be applied
widely in automotive industries. Rao et al. [103] obtained a bimodal ferrite grain structure
through warm rolling and subsequent intercritical annealing of a vanadium-niobium alloyed

18
steel. In the TMP route, the hot-rolled ferrite-pearlite steel was initially normalised at 930 ℃ for
30 min to remove the banded microstructure. The normalised steel was warm rolled to a true
strain of 2.4 (91% reduction in thickness) in the temperature range of 450 to 550 ℃ then air
cooled to room temperature. The warm-rolled samples were finally subjected to intercritical
annealing at 760 ℃ for 2 min followed by water quenching to obtain the DP microstructure. The
bimodal ferrite grain size distribution consisted of both fine ferrite grains ranging from 0.5-2 μm
and relatively coarse ferrite grains ranging from 2-6 μm, as shown in Fig. 12. This TMP-treated
steel exhibited the highest strength and ductility combination for ultrafine-grained DP steels
reported so far in the literature, with UTS of 1371 MPa and uniform elongation of 16%. This
TMP method provides a new path for developing bimodal grain size distribution with the aim of
obtaining an optimised combination of strength and uniform elongation of ultrafine-grained DP
steels.

Fig. 12. Bimodal ferrite grain size distribution of a vanadium-niobium alloyed steel
after warm rolling and subsequent intercritical annealing treatment [103].

In industrial production of DP steels, ultra-fast cooling is usually used in TMP-controlled


rolling to achieve the required mechanical properties. The critical points of ultra-fast cooling
include: (i) large deformation and strain accumulation in the austenite region to induce
hardening of austenite; (ii) ultra-fast cooling operation after rolling to keep austenite in a
hardened state; (iii) stopping of the cooling process at the austenite-to-ferrite phase
transformation point; and (iv) subsequent cooling route control to obtain the desired
microstructure and target mechanical properties [104]. Ultra-fast cooling can realise cooling
rates in the range of 250-500 ℃/s for hot strip with thickness values between 6 and 3 mm [105].
For example, the cooling rate used in ultra-fast cooling process can reach 300 ℃/s which
corresponds to the heat transfer rate of 4.37 MW for a 4 mm-thick hot strip [106-108]. Cai et al.
[109] studied the effects of ultra-fast cooling on the microstructure and mechanical properties of

19
a DP steel under laboratory and industrial production line conditions. In this approach, ultra-fast
cooling was positioned after hot rolling, following by air cooling and laminar cooling, and
finally coiling, as schematically shown in Fig. 13. Using ultra-fast cooling, a volume fraction of
10-12% lath-type martensite with an average ferrite grain size of 6 μm, along with a tensile
strength higher than 590 MPa and YS more than 400 MPa and elongation greater than 22% was
obtained. As proposed by Cai et al. [109], the phase strengthening rather than grain refinement
was realised, and the ferrite grain size obtained was therefore coarser than that obtained by early
positioned ultra-fast cooling and short-interval multi-pass hot rolling in stable austenite region
[110]. Ultra-fast cooling offers a cooling process with flexible paths and can be easily
implemented in practical production line conditions. Therefore, the application of ultra-fast
cooling in TMP-controlled rolling of DP steels is of practical significance and deserves further
research.

Fig. 13. Application of ultra-fast cooling in the TMP-controlled rolling of DP steel [109].
UFC: ultra-fast cooling.

2.2 Thermomechancal processing of TRIP steels


TRIP steels have a microstructure consisting of retained austenite embedded in a ferrite
matrix. The minimum volume fraction of retained austenite in TRIP steels is 5%. Other hard
phases, including martensite and bainite, are also present in varying amounts. Fig. 14 shows a
schematic of the microstructure of TRIP steels [42]. The dispersion of a hard second phase in a
soft ferrite matrix causes a high rate of work hardening during deformation. With increasing
strain, the metastable retained austenite is progressively transformed to martensite, thus
increasing the strength by the phenomenon of strain hardening, in addition to the dispersal of
hard phases. Strain hardening in TRIP steels is due to the continuous addition of hard martensite
to the microstructure, along with the generation of dislocations and internal stresses in the

20
surrounding phases [111]. Such unique transformation allows for a high hardening rate at very
high strain levels, and contributes to enhanced ductility—known as the “TRIP” effect. Yi [112]
reviewed the concept and current progress in TRIP steels, focusing on the methodology for
retention of δ-ferrite in casting, rolling and welding conditions, microstructure evolution by
austempering, as well as the microstructure-property relationship involving the roles of blocky
and lath retained austenite. Li et al. [113] reviewed the progress in TRIP steels from the 1980s
to 2003 with a brief introduction on the requirements of chemical compostions, stabilisation of
retained austenite, composition design, microstructure control and mechanical properties.

Ferrite
Martensite

Bainite

Retained
austenite

Fig. 14. Schematic microstructure of TRIP steels [42].

TRIP steels are generally processed via a two-step heat treatment of intercritical annealing
plus austempering in the bainitic transformation temperature region, followed by quenching.
Such processing results in a final microstructure of equiaxed ferrite surrounded by discrete
particles of martensite/retained austenite and bainite [111]. TRIP steels are promising materials
for automotive applications owing to the unique combination of high strength and high
formability. They are replacing DP and HSLA steels because they have excellent crash energy
absorption due to high work hardening during crash deformation and the ability to be formed
into complex components with deep drawing. Fig. 15 compares the engineering stress-strain
behaviour of TRIP, DP and HSLA steels with similar YS [11]. It can be seen that the TRIP steel
has a lower initial work-hardening rate than the DP steel, but the hardening rate persists at
higher strains where work hardening of the DP and HSLA steels begins to diminish.

21
Fig. 15. Engineering stress-strain curves of TRIP, DP and HSLA steels with a similar YS [11].

The excellent mechanical properties of TRIP steels come from the strain-induced
transformation of retained austenite to martensite at room temperature, which may lead to
redistribution of stresses and a composite effect during deformation [114]. A good balance
between strength and ductility can be achieved if the TRIP effect can be triggered at an
appropriate level by optimal stabilisation of retained austenite. The amount of retained austenite
is a key factor in controlling the final mechanical properties of TRIP steels. It has been shown
that an increase in the volume fraction of retained austenite increases the strain-hardening
coefficient. However, a large amount of retained austenite does not necessarily induce a higher
uniform elongation because sufficient carbon content is required to stabilise the large amount of
retained austenite [115]. If the retained austenite undergoes strain-induced martensitic
transformation at very low strains, or is very resistant to such transformation, the TRIP effect
will be minimal. The amount of strain to initiate the TRIP phenomenon is affected by the
stability of retained austenite. The higher the stability of retained austenite, the later austenite-
to-martensite transformation is delayed during deformation. A TRIP steel with highly stabilised
retained austenite can allow austenite to remain until a failure or crash transforms it into
martensite, which improves the formability and safety when steel is used for automotive parts.
Therefore, the volume fraction of retained austenite and its stability must be carefully controlled
in order to obtain the required combination of strength and ductility of TRIP steels.
Many factors—including carbon content, retained austenite grain size, retained austenite
morphology, phase surrounded retained austenite and steel’s chemical compositions—affect the
stability of retained austenite. The main parameter that represents the stability of retained
austenite is the martensitic transformation starting temperature Ms. The lower the temperature
Ms, the more stable the austenite. The supersaturation of carbon in retained austenite determines
the chemical stability, whereas the size and morphology of retained austenite controls the
mechanical stability [116,117]. Chemical stability of retained austenite due to the enrichment of
carbon is the most important operational mechanism; and carbon content is the main factor

22
determining the chemical stability of retained austenite at room temperature. The carbon content
determines the chemical driving force for the transformation of retained austenite to martensite,
the stress-free transformation strain, and the flow behaviour of retained austenite [118]. Higher
carbon content induces an increased volume fraction of retained austenite in the final
microstructure. Ms will decrease when the amount of carbon enriched in retained austenite
increases [119]. It has been reported that carbon content of about 1% or higher is required to
stabilise the retained austenite [120]. The carbon content of retained austenite must be high
enough, and should be in an appropriate range, to ensure that the retained austenite is not only
thermally but also mechanically stable [121]. TRIP steels use higher quantities of carbon than
DP steels to obtain sufficient carbon content for stabilising the retained austenite phase at below
the ambient temperature. Carbon content in TRIP steels is generally limited to 0.20-0.25 wt.%
due to weldability concerns [122].
Manganese is an austenite stabiliser and it promotes the generation of an adequate amount of
retained austenite and improves the mechanical properties of TRIP steels. However, the content
of manganese should be carefully controlled to ensure TRIP steels have a good coatability and
hardenability. The manganese content in TRIP steels is usually controlled at around 1.5 wt.%
[123]. In TRIP steels, silicon is usually regarded as an important alloying element as it can
promote the formation of carbon supersaturated retained austenite, instead of carbide
precipitation, due to low silicon solubility in carbide [124,125]. However, the presence of
silicon in these steels is associated with some industrial disadvantages, such as poor surface
quality of hot-rolled plates and low weldability [126]. A high content of silicon in TRIP steels
can cause the formation of a stable oxide layer, which will prevent the formation of the
inhibition layer during the hot dip galvanising process [63]. A high level of silicon content,
above 0.5 wt.%, will hinder coating [123,127]. Consequently, there is a trend to reduce silicon
content through process optimisation and/or by partially or completely substituting silicon with
alternative alloying elements such as aluminium.
Aluminium is the best alternative element to substitute silicon in TRIP steels [128-131].
Aluminium is insoluble in cementite, decreases the carbon activity coefficient and increases the
solubility of carbon in ferrite, leading to the higher enrichment in carbon of retained austenite
[132]. It retards cementite formation during bainitic transformation and benefits: the retention of
austenite at room temperature; and the optimisation of the TRIP effect during straining [133]. In
addition, aluminum can accelerate the formation of bainite, which is important for the
production of TRIP steels in continuous galvanising lines, owing to the limited processing time
[117]. C-Mn-Al-type TRIP steels have been a direction of further development of AHSSs [134].
However, aluminium increases the Ms temperature, potentially to above room temperature,

23
reducing the stability of the retained austenite [135]. A full substitution of silicon by an
equivalent amount of aluminium would cause a poorer strength and ductility balance, and mixed
Al-Si TRIP steels are recommended. Other kinds of alloying elements, including molybdenum,
niobium, titanium, chromium, nickel and copper, are also added separately or in combination in
TRIP steels to optimise microstructure and improve mechanical properties [111,126,134,136-
140].
For a given TRIP steel, the other way to improve the stability of retained austenite is to
refine its grain size to an appropriate level. The stability of retained austenite increases as the
grain size decreases. The work of Wang and Zwaag [119] indicated that a decrease in retained
austenite grain size induced a significant decrease in the Ms temperature. Retained austenite
with a grain size smaller than 0.01 μm is useless for TRIP steels as it will not transform to
martensite. A grain size larger than 1 μm may be equally useless as it will immediately
transform to martensite upon cooling or during application of low stress. High stability of
retained austenite with a fine grain size can be achieved after intercritical annealing due to the
absence of substructures in the austenite e.g. stacking faults and other defects that provide
nucleation sites for martensitic transformation [141]. In general, retained austenite is more
stable during deformation when it exists as films between bainitic ferrite plates, rather than
blocky grains between sheaves of bainitic ferrite [142]. The phase-surrounded retained austenite
is also important for its stabilisation. If the retained austenite is surrounded by ferrite, stress will
be transferred to retained austenite from the first deformed ferrite during deformation. Retained
austenite deforms and is transformed due to the TRIP effect. When the retained austenite is
surrounded by hard phase, such as bainite, more stress is partitioned to the bainite during
deformation and retained austenite transformation progresses slowly and possibly not at all
[111].
In order to achieve the best combination of mechanical strength and ductility of TRIP steels,
other microstructural constituents, including ferrite, bainite and/or martensite, should also be
properly controlled in addition to the volume fraction and stability of retained austenite. A TRIP
steel with adequate retained austenite and optimal multiphase microstructural combination is
essential for achieving the best overall performance. To obtain the multiphase microstructure of
TRIP steels, a mixture of austenite and ferrite with nearly the same volume fraction of about
50% is usually achieved before the isothermal bainitic treatment [143]. For conventional hot
rolling processing of TRIP steels, the control of such mixed microstructure is commonly
realised by adjusting the cooling rate and time through multi-stage cooling, after hot rolling of
austenite [144,145]. For obtaining an optimal multiphase microstructure, cold-rolled TRIP steels
are usually processed with a two-step heat treatment i.e. intercritical annealing and isothermal

24
bainite treatment [146]. A schematic illustration of the heat treatment processing of cold-rolled
TRIP steels is shown in Fig. 16. During intercritical annealing, a mixed microstructure of
austenite and ferrite is formed, and carbon atoms are injected into the formed austenite phase for
the first time. Such carbon enrichment, however, is usually insufficient to ensure austenite
remains at room temperature. The isothermal bainite treatment acts as the second carbon
enrichment process through which a quantity of austenite with a high carbon content, where the
martensite transformation temperature Ms is below room temperature, will remain in the final
microstructure after cooling to room temperature. The isothermal bainite treatment has a
significant influence on the microstructures, and thus the mechanical behaviour of TRIP steels.
The heat treatment technologies can normally yield the best combination of strength-ductility in
TRIP steels. However, the major drawback of heat treatment technologies is the difficulty in
precise realisation in industry applications, limiting mass production of TRIP steels.

Ac3 A
Intercritical annealing F
Ac1

Ferrite
Pearlite
Bainite
Isothermal bainite treatment
Ms
Mf
M
F
A
Cold rolling B

Fig. 16. Schematic diagram of heat treatment processing of cold-rolled


TRIP steels. A: austenite, B: bainite, F: ferrite, M: martensite.

As an alternative and effective processing technology, TMP is thought to be the core of


many investigations owing to its potential energy saving and higher productivity—it can
minimise or even eliminate the need for further heat treatment [147-150]. The primary aim of
TMP is to refine the ferrite grain size by the control of the austenite microstructure at the point
of transformation [151]. Ferrite grain refinement is mainly achieved through austenite-to-ferrite
transformation, however the ferrite grain refinement is limited if the ferrite is produced from
recrystallised strain-free austenite. The accumulation of retained strain in unrecrystallised
austenite affects the transformation to ferrite by increasing the number of possible nucleation
sites for ferrite grains. This leads to a progressive reduction of the ferrite grain size after

25
transformation during cooling to room temperature, with increased deformation in the
unrecrystallised region. The cooling rate and original austenite grain size also affect the ferrite
grain size [148,151]. In TRIP steels, another objective of TMP is to refine the bainite
microstructure by deformation in the unrecrystallised austenite region due to the retarding effect
of “pancaking” [151]. The bainite microstructure can be significantly improved by deformation
of more than 50% in the unrecrystallised region [152]. Also, TMP has an influence on the
transformation behaviour, which may lead to different morphologies of the second phase.
Acicular ferrite has been found to induce a significant increase in the quality and stability of
retained austenite, which improves the room temperature mechanical properties of TRIP steels
[152]. Deformation in the unrecrystallised region introduces defects to the matrix microstructure,
which can affect the stability of retained austenite and influence the mechanical properties of
TRIP steels [153]. Deformation in the unrecrystallised region can change the amount of retained
austenite and reach the optimal TRIP effect. This is mainly because the driving force for
diffusion transformation is increased and the shear transformation is delayed by introducing the
defects. A typical TMP procedure consists of three stages [154]: (i) deformation in the
austenite-recrystallisation region; (ii) deformation in the austenite-unrecrystallisation region;
and (iii) deformation in the “austenite + ferrite” (intercritical) region. In stage one, coarse
austenite grains are refined by repeated deformation and recrystallisation; in stage two,
deformation bands and intragranular lattice defects in elongated unrecrystallised austenite are
formed; and in stage three, the second stage continues, as well as the continuous formation of
subgrains within deformed ferrite grains. However, as the microstructure and final mechanical
properties are very sensitive to the processing schedules, the final microstructure—in terms of
retained austenite and other microstructure constituents—must be carefully and precisely
controlled through process optimisation. Moreover, the chemical compositions of a TRIP steel
have an important effect on optimal processing parameters, and this should be taken into
account in the design of TMP schedules.
The work of Hosseini et al. [155] showed that the state of prior austenite greatly influenced
the stability of retained austenite, as well as the final phase combinations. Refinement of parent
austenite leads to a smaller bainite pocket size with thin interlayer spaces after TMP, which
enhances the stability of retained austenite through a mechanism of mechanical stabilisation. An
increase in the prior austenite grain size will result in a microstructure consisting of less retained
austenite and martensite/austenite after TMP treatment, which in turn induces reduced
mechanical properties. To study the influence of austenite deformation on microstructural
development in TRIP steel, Hosseini et al. [154] adopted a multi-stage isothermal
deformation/cooling TMP program to deform the steel with different strains at various

26
temperatures in the unrecrystallised austenite region, as shown in Fig. 17. They found that the
maximum volume fraction and carbon content of retained austenite could be achieved by
deforming the steel to some intermediate strains in the unrecrystallised region. However, further
straining would result in a drastic reduction of both the retained austenite amount and the carbon
content in it due to the formation of pearlite. Pearlite acts like a carbon sink and thus decreases
retained austenite stability. It has been reported that when the reduction is approximately 30%,
the formation of serrated austenite grain boundaries is considered to be the most probable reason
for the increase in the nucleation rate of pearlite on the boundaries [156]. Moreover, a decrease
in deformation temperature in the unrecrystallised region had a beneficial effect on the increase
of both the volume fraction and carbon content of retained austenite.

Fig. 17. Schematic diagram of TMP schedule of a low-silicon TRIP steel with different
strains at various temperatures in the unrecrystallised austenite region [154].

In order to investigate the effects of TMP on the microstructure, morphology and volume
fraction of retained austenite of a low-alloy TRIP steel, Skalova et al. [157] applied seven
processing conditions of intercritical annealing and subsequent heat treatment in the bainitic
region. They found that higher austenitisation resulted in an undesirable coarse and
heterogeneous ferritic-bainitic microstructure. Higher deformation would cause grain elongation
and the presence of lots of microshrinkages in the microstructure, and slower cooling induced
the presence of smaller pearlitic islands in the final structure. The optimal TMP conditions for
the studied TRIP steel consisted of austenitisation at the temperature of 850 ℃ followed by 30%
compression deformation at 700 ℃ and rapid cooling to the bainitic transformation temperature
of 400 ℃. The work of Grajcar [158] indicated that hot working in the austenite+ferrite region
essentially influenced a multiphase structure of TRIP steels. The stability of retained austenite
can be improved when TRIP steels are hot-worked in an “austenite + ferrite” range [159]. Fig.
18 shows a schematic representation of TMP with hot-working in the “austenite + ferrite”
region, in which an essential step is a selection of a suitable cooling rate after hot-working to the

27
holding stage of isothermal bainitic transformation [159]. The cooling rate should be
sufficiently high, i.e. 90 ℃/s by oil cooling, to suppress the formation of pearlite. Both the
temperatures for hot-working and isothermal bainitic transformation have an important effect on
obtaining the required multiphase microstructure. The optimal austenitising temperature should
be beneficial for the carbon concentration in the austenite, and the best temperature of the
bainitic transformation is usually determined by achieving the maximum retained austenite
volume fraction. A low bainitic transformation temperature is generally used to slow the
diffusion of carbon out of the bainite due to the low rate of bainitic transformation. By
processing in the austenite+ferrite region, an increase in the amount of retained austenite can be
achieved through the selection of processing parameters. The stability of retained austenite can
be improved due to the significant decrease in its grain size and enrichment of retained austenite
in carbon during plastic deformation in the “austenite + ferrite” region [159]. The TMP schedule
with hot-working in the “austenite + ferrite” region has a potential for the development of hot
rolling technology of TRIP steels. However, the hot-working and cooling from a finishing
rolling temperature must be precisely controlled to obtain the required stable and high fraction
of retained austenite.

Fig. 18. Schematic representation of TMP of a TRIP-aided microalloyed steel with


hot-working in the “austenite + ferrite” region [159].

Timokhina et al. [148] conducted TMP treatments on two low-carbon TRIP steels with and
without niobium additions. The TMP schedules with continuous and discontinuous cooling
methods are illustrated in Fig. 19a and b, respectively [148]. In both schedules, deformation at
1100 ℃ was in the austenite recrystallised region. Further deformation at 850 ℃ for the niobium
steel and 825 ℃ for the niobium-free steel was in the austenite-unrecrystallisation region. The 2
min holding after 25% deformation produced a uniform recrystallised austenite microstructure,
and further deformation at 850 ℃ for the niobium-containing steel and 825 ℃ for the niobium-

28
free steel in the unrecrystallisation region refined the ferrite grain size and increased the ferrite
and bainite transformation temperatures. The results indicated that the addition of niobium
promoted the formation of acicular ferrite and stable retained austenite, and the chemical
compositions affected the TMP schedules. Optimal mechanical properties of the niobium-
containing TRIP steel were obtained after isothermal holding at 400 ℃, as this treatment could
induce the maximum volume fraction of retained austenite with acicular ferrite as the
predominant second phase. Both the amount and the morphology and stability of retained
austenite have an important impact on the optimal mechanical properties. When deformed in the
unrecrystallisation region, a decrease in deformation temperature increases the volume fraction
of retained austenite in the final microstructure [160]. This behaviour is thought to be
attributable to the slower kinetics of restoration processes at lower deformation temperatures,
which leads to more strain energy being accumulated in microstructure. The fraction of retained
austenite can reach the maximum level at an intermediate strain and, after a plateau, decreases.
The initial increase in the fraction of retained austenite with straining is ascribed to a number of
factors, such as increasing the amount of strain energy accumulated, the development of
subgrains in ferrite grains, and increasing the density of ferrite nucleation sites. The subsequent
reduction of retained austenite fraction at a higher amount of strains is also attributed to the
pearlite formation at serrated austenite grain boundaries induced by deformation [154].

(a) (b)

Fig. 19. TMP schedules of TRIP steels with (a) continuous cooling and (b) discontinuous
cooling [148]. Note: TC represents simulated coiling temperature (T C = 550 ℃, 500 ℃, 450 ℃,
400 ℃ and 350 ℃).

Zrnik et al. [161,162] conducted a systematic investigation of the effects of hot deformation
introduced in austenite recrystallisation, unrecrystallisation and intercritical temperature regions
on the austenite conditioning and microstructural evolution of a TRIP steel. It was found that
the choice of applied strain and temperature of deformation prior to transformation had a strong
impact on transformation kinetics, final multiphase (ferrite, bainite and retained austenite)

29
structure characteristics and mechanical properties of TRIP steel. A higher volume fraction of
retained austenite could be obtained when two successive compressive deformations were
inserted at austenite conditioning and when the second straining was applied at intercritical
temperature. An increased retained austenite volume fraction increases the ductility markedly
but may reduce the strength of TRIP steel to a small degree. Muransky et al. [163] compared the
mechanical properties of a low-alloyed TRIP steel processed by different TMP schedules
according to Zrnik et al. [161], and found that the microstructure with lower retained austenite
volume fraction and finer ferrite matrix exhibited higher YS but lower tensile strength and
elongation than the microstructure consisting of a higher retained austenite volume fraction and
coarser polygonal ferrite matrix. The difference in mechanical properties is caused by the
differences in mechanical properties of the constituent phases, which may lead to different load
partitioning during the collaborative TRIP deformation mechanism.
In order to obtain ultrafine-grained ferrite in the matrix of a low-carbon TRIP steel, Chen et al.
[164] combined DIFT and quenching and portioning (Q&P) processes through TMP treatment, as
schematically illustrated in Fig. 20. The highest UTS of about 1603 MPa could be achieved by
DIFT process solely without partition at 300 °C. However, its elongation was the lowest, at about
11%. Interestingly, the UTS of the TRIP steel treated by the combined DIFT+Q&P process could
reach to be higher than 1200 MPa with the elongation above 16% due to the introduction of fine-
grained ferrite (about 5 μm in diameter) with lath martensite and retained austenite. Based on their
research findings, it is possible to further improve strength and ductility by optimising the
processing parameters. Wakita et al. [165] applied an advanced TMP to obtain ultrafine ferrite
microstructure in a low-carbon TRIP-aided multiphase steel. Through heavy deformation in a
supercooled austenite region, high volume fraction of retained austenite and ultrafine ferrite with
a grain size of 2 μm could be achieved. The TMP-treated TRIP steel with ultrafine microstructure
showed a good combination of tensile strength (1080 MPa) and elongation (26.9%). Similarly,
Sun et al. [166] and Yang et al. [167,168] developed a TMP technology based on dynamic
transformation of undercooled austenite. Fig. 21 shows a typical schematic diagram of the TMP
of two TRIP steels with and without niobium additions based on dynamic transformation of
undercooled austenite [169]. The key feature of this process is that the formation of multiple-
phase mixed microstructure can be well controlled by the applied strain of hot deformation of the
undercooled austenite. In comparison with hot-rolled TRIP steel by the conventional controlled-
cooling processing, the multiphase microstructure of hot-rolled TRIP steels based on dynamic
transformation of undercooled austenite processing could be easily controlled and markedly
refined, demonstrating combinations of strength and ductility [143]. For instance, the mechanical
properties of the C–Mn–Al–Si TRIP steel based on dynamic transformation of undercooled

30
austenite processing can reach 780 MPa and 34% for the tensile strength and total elongation
respectively, which is nearly the same as that of the cold-rolled TRIP steel of similar composition
[168]. Also, compared to controlling the cooling rate, strain is easy to control precisely in industry
[170].

Fig. 20. TMP schedule combined of DIFT and Q&P processes [164].

Fig. 21. Schematic diagram of TMP of two TRIP steels with and without niobium-additions
based on dynamic transformation of undercooled austenite [169].

Apart from the aforementioned TMP technologies, several other TMP schedules have been
recently developed for the optimisation of microstructure and improvement of mechanical
properties of TRIP steels. For example, Ranjan et al. [171] applied three TMP routes to three
TRIP steels with different compositions based on available literature and thermodynamics-
based calculations. The details of the three TMP schedules, along with a schematic
representation of the changes in the microstructures, are shown in Fig. 22. These TMP
schedules involve: (i) a small isothermal step in the intercritical temperature region to obtain
50% ferrite and enrich the austenite with carbon; and (ii) an isothermal holding in the bainite

31
region to obtain bainite and further enrich the austenite with carbon to enhance its stability.
Based on their research findings, the amount of retained austenite and its carbon content of a
given TRIP steel can be greatly increased when the TMP cycle does not involve any isothermal
treatment in the intercritical temperature range. Guo et al. [172,173] developed a TMP schedule
for high-manganese TRIP steels based on warm deformation of martensite and subsequent
short-time annealing in the intercritical region. This method solved the problem of producing
high-manganese TRIP steels with the microstructure consisting of ultrafine-grained ferrite
matrix and martensite/retained austenite islands. In this method, high-manganese TRIP steels
are commonly heat-treated by long-time intercritical annealing in the ferrite-austenite two-phase
region, or in the intercritical region directly after hot rolling of austenite or after hot rolling of
austenite plus cold rolling of martensite [174-176]. The mechanical properties of high-
manganese TRIP steels treated by this developed TMP technology can be greatly improved to
be comparable to the steels with similar compositions subjected to intercritical annealing for
hours after cold rolling of martensite, indicating the effectiveness for time and cost savings,
which are significant for mass production in industry. In addition, ultra-fast heat treatment was
suggested in the TMP of TRIP steels [177]. Ultra-fast heat treatment limits macro-diffusion, and
can affect phase transformations and contribute to grain refinement by suppression of
recrystallisation. Ultra-fast heat treatment has been regarded as a new opportunity for the design
of AHSSs with heterogeneous-phase microstructures [178]. However, ultra-fast heat treatment
requires a specified device to realise ultra-fast reheating cycles, and has difficulties in
processing large-sized steels, which will limit its applications in practical production lines of
AHSSs.

32
Fig. 22. Schematic representation of three TMP schedules of TRIP steels [171].

2.3 Thermomechancal processing of CP steels


The microstructure of CP steels is composed of ferrite/bainite matrix with a small amount of
retained austenite, martensite and pearlite. Generally, CP steels have a carbon content level
lower than 0.15 wt.%, and the alloying elements used are very similar to that of DP and TRIP
steels. Additionally, CP steels contain small quantities of niobium, titanium and/or vanadium to
form fine precipitates and realise the precipitation strengthening effect [4,122,179]. An extreme
grain refinement can be achieved by retarded recrystallisation or precipitation of microalloying
elements such as titanium or niobium. The volume fractions of phases contained in CP steels
affect the mechanical properties and vary with the cooling rate from the austenite region. A
wide range of properties can be obtained with the same chemical compositions only by
adjusting the volume fractions of the phases in CP steels through the control of the cooling rate.
The cooling rate influences the formation of ferrite and bainite/tempered martensite, but has no
impact on the amount of retained austenite. High cooling rates suppress ferritic formation and
favour the martensitic transformation [180]. The dependence of phase fractions on the cooling
rate from austenite region of a CP steel is illustrated in Fig. 23 [180].

33
Phase fraction (%)

Cooling rate (K/s)


Fig. 23. Dependence of phase fraction on the cooling rate of a CP 800 steel [180].

CP steels are strengthened by solid solution, precipitation, grain refinement and phase
transformation mechanisms. In comparison to DP steels, CP steels have shown much higher YS
at equal tensile strength levels of 800 MPa and greater [42]. In deformation, CP steels exhibit
continuous yielding behaviour, and are characterised by high energy absorption, high residual
deformation capacity and good hole expansion. Fig. 24 shows the engineering stress-strain
curves of some CP steel grades compared to that of mild steel [11]. It can be seen that CP steels
show significantly higher strength levels and much lower strains than that of mild steel. The
ductility of CP steels exhibits a decreasing trend from 13% to 3% with an increase of the UTS
from 850 to 1200 MPa.

1400
Engineering stress (MPa)

1200
CP 1000/1200
1000
CP 800/1000
800
CP 650/850
600
400 Mild Steel
200
0
0 5 10 15 20 25 30 35 40 45 50
Engineering strain (%)

Fig. 24. Engineering stress-strain curves of some CP steel grades and a mild steel [11].

Titanium nitride TiN is the most stable precipitate and is known to have a dissolution
temperature higher than the iron liquidus temperature [64]. Grajcar [181] conducted a
thermodynamic analysis of precipitation processes in austenite of a titanium-niobium
microalloyed steel, and found that TiN possesses the lowest solubility product and therefore
retains the highest stability compared to other precipitates, including NbC, Nb(C,N) and TiC.

34
TiN is the first phase to precipitate in austenite during steel cooling [182]. Based on the
thermodynamic model proposed by Grajcar [181], precipitation of TiN takes place at about
1500 ℃, and with a decrease in temperature, a gradual decrease in the content of nitrogen and
titanium dissolved in solid solution and an increase in the amount of TiN will occur. The
finishing precipitation temperature is about 1050 ℃ which corresponds to a complete bounding
of titanium with nitrogen as TiN in austenite. NbC is the next phase to precipitate in steel with a
continuous decrease in temperature, with a starting precipitation temperature of about 1280 ℃
and finishing precipitation temperature of about 900 ℃, as indicated in Fig. 25 [181]. It should
be noted that in steels containing both titanium and niobium, precipitation of NbC will be
delayed due to the formation of stable (Nb,Ti)(C,N) [183]. As a sequence of precipitation with
the decrease in temperature, TiC is the next phase to precipitate in steel. TiC has a starting
precipitation temperature of about 1080 ℃ which may vary depending on the change of real
titanium content required to bound the whole nitrogen in steel [181].

1600
1500
1400 TiN
Temperature (℃)

1300
1200 NbC + TiN

1100
TiC + NbC + TiN
1000
900
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Mass fraction of precipitate
(%)
Fig. 25. Temperature sequence of precipitation of MX-type interstitial phases in austenite [181].

Vanadium is the most soluble of the three elements titanium, niobium and vanadium, and it
does not readily precipitate in austenite. When the steel is heated, VC initially dissolves in
ferrite, followed by VN at approximately 700 ℃ [184]. The solubility of VC is considerably
higher than those of other microalloying carbides and nitrides, and it will completely dissolve
even at low austenitising temperature. As a result, vanadium starts to precipitate in austenite as
almost pure vanadium nitride of VN during cooling until all the nitrogen is consumed [185].
There is a gradual transition to form mixed vanadium carbonitrides V(C,N) when the nitrogen is
about to be exhausted, and nitrogen-rich V(C,N) may precipitate during the following austenite-
to-ferrite transformation. Since V(C,N) has high solubility in austenite, it hardly affects the
austenite grain size during hot working but can precipitate during the subsequent cooling to
increase the strength level of steel by precipitation strengthening. A combination of vanadium

35
and microalloying elements titanium and niobium in steel usually leads to the formation of
precipitates with complex chemical compositions [186,187]. For example, the initially formed
TiN may incorporate a significant amount of niobium and vanadium outside its layer to be a
complex precipitate. In fact, because titanium, niobium and vanadium carbides, nitrides or
carbonitrides have similar crystal structures, co-precipitation is very common in multi-
microalloyed steels [64]. With an in-depth understanding of the precipitation behaviour of
various precipitates, it is possible to determine the most favourable combination of TMP
parameters which can achieve the optimal final microstructure with required mechanical
properties. Niobium alloying in CP steels is very popular as it can refine the microstructure,
facilitate the formation of acicular ferrite and increase the stability of retained austenite, in
which carbon content and volume fraction are not affected apparently. As a result, the strength
of niobium-containing CP steels can be enhanced significantly without deterioration of
elongation.
The roles of microalloying elements such as titanium, niobium and vanadium in CP steels
include: (i) adjusting the state of austenite prior to the transformation of austenite-to-ferrite, thus
accelerating the formation of fine ferrite grains [170]; (ii) enhancing the stability of retained
austenite by increasing the stacking fault energy of retained austenite due to the dissolved
microalloying elements in the final complex microstructure [169]; and (iii) achieving solid
solution and precipitation strengthening [144]. Among these roles, grain refinement is the key
target in the chemical compositions design of CP steels by microalloying additions. Generally, the
aim of the TMP of CP steels is to produce microstructures containing 50% of polygonal ferrite
and the maximum achievable volume fraction of retained austenite, with the remainder being
carbide-free bainite [117,188]. Fu et al. [189] thought the final microstructure of niobium-
containing CP steels should be consisted of 50%-60% ferrite matrix and the maximum volume
fraction of retained austenite. To achieve this, they applied a TMP on a 0.038 wt.% niobium
microalloyed CP steel based on the dynamic transformation of undercooled austenite, as shown
in Fig. 26 [189]. In the TMP, CP steel was first austenitised at T1 temperature (1250 ℃) for 5
min, after that the steel was cooled at 5 ℃/s to T2 temperature (1100 ℃) followed by
deformation at a strain rate of 1 s-1 to 0.36 plastic strain, then cooled at 5 ℃/s to T3 temperature
(780 ℃) where the austenite was undercooled. At this temperature, the steel was deformed at 1
s-1 to a strain of 0.69, and finally cooled at 30 ℃/s to 450 ℃ and held for 3 min followed by
water quenching to room temperature. The processed CP steel consists of 50%-60% of equiaxed
ferrite and acicular ferrite (~35%), 8.3% retained austenite and the remainder of martensite and
bainite. This microstructure can lead to good mechanical properties with YS of 654 MPa, UTS
of 845 MPa and total elongation of 34%.

36
Fig. 26. Schematic diagram of TMP based on dynamic transformation of undercooled austenite [189].

In order to ensure microalloying elements are efficiently utilised in CP steels, a careful design
of processing technologies becomes essential, with the aim of achieving optimal microstructure
and enhanced mechanical properties. TMP is an energy-saving and effective technology for
processing CP steels, and the desired microstructural constituents can be easily obtained through
application of appropriate processing parameters. One key step in the design of TMP strategies of
CP steels is to first understand the precipitation behaviour and stability of various precipitates as
they have significant effects on the decomposition of austenite during processing treatment, and
thus the refinement of final microstructure. The kinetics of austenite decomposition in CP steels is
governed by the steel’s chemical compositions and the austenite grain size, as well as the amount,
state and stability of various precipitates. These precipitates include carbides, nitrides and
carbonitrides. The presence of precipitates increases the austenite coarsening temperature and
effectively retards austenite grain growth [190], which is of great significance in the selection of
reheating temperature in hot working. Knowledge of the evolution of precipitates with
temperature is essential in determining the most favourable combination of heat treatment
parameters.

2.4 Thermomechancal processing of MART steels


MART steels are, as the name suggests, mostly martensitic with, in some cases, a small
amount of ferrite and/or bainite. The strength, hardness and hardenability of MART steels
increase by increasing the carbon content. Other alloying elements such as silicon, manganese,
chromium, nickel, molybdenum, boron and vanadium are also used in various combinations to
increase hardenability. The resulting MART steels are best known for their extremely high
strength, with UTS up to 1700 MPa or even higher [42,191]. A tensile strength comparison
between MART and mild steels is illustrated in Fig. 27 [11]. Obviously, MART steels show
37
significantly higher strength than that of mild steel. MART steels are often subjected to post-
quench tempering to improve the ductility and provide adequate formability even at extremely
high strength levels. MART steels are typically roll formed and used where high strength is
critical. They may be bake hardened and electrogalvanised for applications requiring corrosion
resistance.

1600
MART 1150/1400
1400
Engineering stress (MPa)

1200
MART 950/1200
1000
800
600
400 Mild steel
200
0
0 5 10 15 20 25 30 35 40 45 50
Engineering strain (%)
Fig. 27. Engineering stress-strain curves of two MART steels and a mild steel [11].

Martensite is the hardest microstructural constituent of steels, and the strength of MART
steels is very high to allow use in advanced car body construction, especially in areas where
crash loads are expected. However, one of the concerns about MART steels is their sensitivity
to cracking, particularly under impact and bending deformation, in crash situations. Another
problem is delayed cracking observed in ultra-high strength MART steels [192]. Optimisation
of the microstructure in an appropriate way can help to reduce sensitivity to cracking. For body-
centred cubic (BCC) martensite, failure generally occurs by transcrystalline cleavage fracture on
the {100} lattice plane [193]. In order to enhance the cracking resistance, high angle grain
boundaries with neighbouring grains having different orientations to the acting major stress can
be introduced, which can be achieved through grain refinement [194]. Grain size effects in lath
martensitic steels have been deeply investigated and a relationship between the microstructure
and mechanical properties has been established [195]. Extensive research has shown that with a
decrease in grain size, the sizes of both the packet (laths with the same habit plane) and block
(laths of two specific variants group) decrease [196-200]. The structure of typical lath
martensite can be generally described by the descending levels of the prior austenite grain,
packet, block and lath, according to their crystallographic features [201,202]. In martensite, a
prior austenite grain, which contains a large number of discrete laths of dislocated structure, is
divided into several martensitic packets in which the laths share the same habit plane. Each
packet is further divided into martensitic blocks in which the parallel laths possess the same

38
orientation. In the typical microstructure of dislocated martensite, martensite laths are oriented
in a Kurdjumov-Sachs relationship in which they share a close-packed plane and close-packed
direction with the parent austenite ((111)γ//(011)α') [203]. The lath boundaries are low angle
boundaries that do not impose significant crystallographic discontinuities. With regard to the
dependence of mechanical properties on the structure of martensite, the relevant effective grain
size should be considered in Hall-Petch type relationships. It has been widely accepted that the
packet size can be regarded as the effective grain size of martensite [204-206], and the block is
usually thought to be the minimum structure unit controlling strength and toughness of
martensite [207].
The martensitic microstructure in MART steels typically forms during rapid cooling, such as
quenching, on the run-out table or in the cooling section of the continuous line following hot
rolling, annealing, or a post-forming heat treatment in which austenite exists. The austenite-to-
martensite transformation is diffusionless, and therefore martensite forms without any
interchange in the position of neighbouring atoms. To produce MART steels, the cooling rate
must be high enough to suppress the decomposition of austenite by diffusional processes to
ferrite, pearlite and/or bainite. The austenite-to-martensite transformation begins when austenite
is cooled to the starting temperature of martensitic transformation, Ms, and completes when the
temperatures drop to the martensitic transformation finishing temperature, Mf. The fraction of
the transformed martensite is generally a function of temperature and it increases continuously
from zero at Ms to completion at Mf [44]. If the cooling stops at a temperature between Ms and
Mf, the transformation will be halted and it may proceed only on further cooling.
For a specific MART steel, TMP provides an effective way for the microstructural
refinement with the purpose of obtaining excellent combination of mechanical properties. In
TMP practice of MART steels, the controlled processing stage plays an important role in
determining the final mechanical properties through adjustment of microstructural constituents.
Bandyopadhyay et al. [208] conducted a typical TMP on a microalloyed low-carbon MART
steel through the control of the rolling stages, and obtained the best combination of strength
(1440 to 1538 MPa) and ductility (11% to 16%). The steel with cross section of 95 × 95 mm2
was at first heated up to 1200 ℃. After holding for 3 h, the steel was rolled with a 42%
reduction in three multiple passes in the temperature range of 1100 to 1020 ℃. Then, the steel
was further rolled with a reduction of 36% in the temperature range of 1020 to 920 ℃. Finally,
it was rolled with a reduction of 54% in three multiple passes in the temperature range of 920 ℃
to different finish rolling temperatures, followed by water quenching. Fine-grained, pancaked
austenite and highly dislocated fine lath martensitic microstructure, along with the presence of
tiny precipitates, are the main causes of the increase in strength of MART steels with the

39
lowering of finish rolling temperature [209]. The work of Kaijalainen et al. [210] indicates that
an increase in the total reduction in the non-recrystallisation region, in conjunction with a
lowering of the finish rolling temperature, promotes the formation of transformation products
such as bainite and ferrite at the expense of lath martensite. Therefore, the total reduction in
rolling should be controlled within an appropriate range in order to obtain full fine lath
martensite after TMP.
Hot rolling temperature has a significant effect on the grain refinement and precipitation
behaviour, and therefore affects the mechanical properties of MART steels. Han et al. [211]
compared two TMP routes on a Ti-microalloyed low-carbon MART steel with different start
rolling temperatures (950 ℃ and 1100 ℃) but with the same reheating temperature (1200 ℃)
and holding time (1 h), as schematically illustrated in Fig. 28. They found that deformation
without recrystallisation (950 ℃) was beneficial for the refinement of both martensitic
microstructure and precipitates. Deformation within the non-recrystallisation region can cause
enough storing deformation energy and increase dislocation density to provide grain nucleation
with driving force and nucleation place, respectively. Furthermore, the grain growth is
suppressed by both the pancaked austenite grain boundaries and the diffused precipitates in the
reheating process. At a relatively low temperature (950 ℃), coarsening of precipitates does not
take place, resulting in the formation of fine-sized precipitates and an improvement in
precipitation hardening [64,212]. Therefore, the steel rolled at a low non-recrystallisation
temperature exhibits a much higher strength than that rolled at a high temperature where
recrystallisation occurs due to the additional dislocation strengthening and enhanced
precipitation strengthening. Asadi and Palkowski [213] simulated the last three deformation
steps of a hot rolling process of two low-carbon MART steels by means of TMP. They indicated
that austenite grains became elongated and deformation bands were introduced when
deformation was applied at temperatures below the non-recrystallisation temperature. The
number of nucleation sites at the austenite grain boundaries, and within the austenite grains,
increases as the amount of deformation in this region increases. As a result, the austenite-to-
martensite transformation from deformed austenite leads to much finer martensite grains than
that from recrystallised strain-free austenite. Therefore, the non-recrystallisation temperature is
a very important parameter and its determination is crucial in the design of controlled TMP
schedules. For a given MART steel, the non-recrystallisation temperature Tnr can be calculated
by the empirical equation proposed by Barbosa et al. [214]:

Tnr = 887C + 4645Nb – 644Nb1/2 + 732V – 230V1/2 + 890Ti + 363Al – 357Si (1)

40
Fig. 28. Schematic diagram of TMP on a Ti-microalloyed low-carbon MART steel [211].

In TMP of MART steels, the state and feature of the parent austenite significantly affect the
transformation behaviour in cooling, which thus influence the final microstructure and
mechanical properties of the processed steels [215]. Hot deformation of austenite promotes the
formation of ferrite and pearlite, but enhances the stability of austenite against bainitic and
martensitic transformations [216,217]. Some research has indicated that the starting temperature
of martensitic transformation (Ms) will decrease when the austenite is deformed due to
mechanical stabilisation [218,219]. However, this effect remains unclear in microscale in terms
of austenitic orientation, austenitic grain size, or eventually the local inhomogeneity of austenite.
Xiao et al. [220] found that the undissolved carbides associated with hot deformation increased
the inhomogeneity of carbon distribution in deformed austenite, and the Ms temperature
increased when the austenite was deformed. The work of Nikravesh et al. [221] indicated that
the deformation of austenite caused increasing nucleation sites of martensite, such as
deformation bands and aggregation sites of dislocations. If deformation temperature is low so
that the strength of austenite after deformation increases, the essential shear for martensitic
transformation will become more difficult, and as a result Ms decreases. If the deformation is
accompanied with DRX, static recrystallisation or full dynamic recovery (DRV), the structure of
austenite will become softened, and so the essential shear for martensitic transformation
becomes easier and Ms increases [221].
Wang et al. [222] conducted research on the effect of hot deformation of austenite on the
martensitic transformation of a high manganese steel by controlled TMP, as schematically
presented in Fig. 29. In high manganese steels, two types of martensite, i.e. ε-martensite
(hexagonal close-packed martensite) and α'-martensite (BCC martensite), generally exist in an
‘austenite → ε-martensite → α'-martensite’ sequence [223]. They found that hot deformation of
austenite (Fig. 29a) hindered martensitic transformation, and α'-martensite was much more

41
effectively restrained than ε-martensite. Orientation, orientation gradients, grain size and
recrystallisation of austenite are four important factors affecting martensitic transformation.
When the deformed austenite grains are fine, the martensitic transformation will be hindered
more effectively, compared to the coarse grains. Large orientation gradients usually function as
an obstacle to the growth of ε-martensite and, combined with grain size, influence the ε-
martensite transformation. After deformation, a strong deformation texture of {110}γ will form,
which supresses the transformation of ε-martensite. Moreover, fine dynamically recrystallised
grains enhance the mechanical stability of austenite and therefore inhibit martensitic
transformation [222]. The recovery treatment (Fig. 29b) will weaken the mechanical
stabilisation of austenite, and thus promote the martensitic transformation.

Fig. 29. Schematic diagrams of TMP of a high manganese steel with (a) the effect of hot
deformation of austenite on martensitic transformation, and (b) the effect of static recovery
and recrystallisation of the deformed austenite on martensitic transformation [222].

In practice, tempering is an important part of the TMP of MART steels in order to improve
their ductility and toughness [224]. The modifications of the mechanical behaviour of MART
steels during conventional tempering are due to microstructural changes, including the carbon
segregation to lattice defects, carbon precipitation, decomposition of RA, and recovery and
recrystallisation of the martensitic microstructure [225-230]. The tempering time and temperature,
the cooling rate from the tempering temperature, and the chemical compositions of the steel—
including carbon content, alloying elements and residual elements—affect the microstructure and
mechanical properties of tempered MART steels. The work of Massardier et al. [231] showed that
rod-like cementite particles precipitated at low tempering temperature, and this precipitation was
gradually replaced by coarser incoherent precipitates as the tempering temperature increases.
Precipitation of rod-like cementite particles causes additional loss of elongation, as well as
toughness [232]. At higher tempering temperature, cementite is spheroidised and ductility, as well
as toughness, will recover. Therefore, carbides could play a key role in the martensite strength in

42
tempering. Massardier et al. [231] also indicated that the carbide precipitation during fast
tempering significantly affected the evolution of the mechanical properties of tempered MART
steels. In general, tempering of MART steels for crash-relevant automotive components in the
temperature range of 200-400 ℃ should be definitely avoided, because the crash property
deteriorates quickly with a tempering treatment in this temperature range. In order to improve the
toughness of MART steels, tempering can be processed in combination with varying quenching
methods. Duan et al. [233] compared the effects of direct quenching and tempering process and
conventional reheat quenching and tempering process on the microstructure and mechanical
properties of a low-carbon MART steel, and found that the tensile strength of the direct-quenched
specimen was higher than that of the reheat-quenched one. The low temperature toughness of the
directly quenched and tempered specimen was generally inferior to that of the reheat quenched
and tempered specimen. Therefore, a strategy should be developed for balancing the strength and
toughness in the tempering process of MART steels.

3. Thermomechanical processing of the second generation AHSSs


3.1 Thermomechancal processing of TWIP steels
TWIP steels have received much interest in recent years due to their excellent combination
of high strength and ductility. TWIP steels are characterised by a high rate of work hardening
resulting from the generation of deformation-induced twins. A large amount of deformation is
driven by the formation of these twins [234]. The twinning induced by this particular
deformation mode causes a high value of instantaneous hardening rate as the microstructure
becomes finer and finer. The twin boundaries act like grain boundaries in their functionality to
strengthen the steels by restricting the movement of dislocations during deformation.
Crystallographic twinning occurs when two separate crystals share some of the same crystal
lattice points in a symmetrical manner. Crystallographic twinning differs from sliding, even
though both have similar macroscopic shear mechanisms that occur at a distinct plane and need
a certain shear stress to be activated. To cause twinning, the shear stress that occurs across the
twinning plane and is resolved in the twinning direction should be positive. An important
difference between deformation twinning and slip is that twinning depends more strongly on
shear direction. Crystallographic twinning is a shear process yielding a significant re-orientation
of the lattice, whereas for crystallographic slip the crystal lattice is nearly not reoriented during
a free shearing deformation process [235]. TWIP steels are austenitic steels with a face-centred
cubic (FCC) crystal structure. The occurrence of crystallographic twinning in TWIP steels is
strongly dependent on the stacking fault energy. Low stacking fault energy (<20 mJ/m2) favours
the transformation of austenite to martensite, i.e. TRIP effect, while high stacking fault energy

43
(>20 mJ/m2) promotes the mechanical TWIP effect [4]. However, the critical stacking fault
energy for achieving TWIP effect is still unclear so far. The critical stacking fault energy may
vary from 18 to 33 mJ/m2 in order to obtain the deformation twins, depending on the chemical
compositions of TWIP steels and the methodologies used for determining stacking fault energy
[236-238]. TWIP steels possess a medium stacking fault energy ranging from 20-40 mJ/m2 at
room temperature, which is beneficial for the generation of twins during deformation [239].
Stacking fault energy is the energy required to change the atomic-layer stacking sequence
during the deformation, and plays a significant role in defining the type of plasticity mechanism
that prevails in high manganese steels. Stacking fault energy is affected by a few decisive
parameters, including chemical compostions, temperature, segregation of alloying elements at
stacking faults, elastic strain, magnetic contribution, and ratio of valence electrons to atoms. In
high manganese steels, the addition of alloying elements can affect the austenite stability and
deformation mechanism through the change of stacking fault energy. Manganese controls the
stacking fault energy level of mechanical twinning in high manganese steels. An increase in
manganese induces increased stacking fault energy, and therefore benefits the occurrence of
TWIP effect [240]. Carbon increases the stacking fault energy and thereby inhibits the
formation of martensite. Aluminium in high manganese steels increases the stacking fault
energy of austenite and suppresses the transformation from austenite to martensite, and
therefore favours the formation of deformation twins [241]. Silicon generally has a negative
effect on stacking fault energy, i.e. an increase in silicon would cause decreased stacking fault
energy [242]. Temperature has a great impact on the stacking fault energy. When the
temperature increases, the stacking fault energy increases as well [243]. The segregation of
alloying elements at stacking faults and magnetic transition will cause a decrease in stacking
fault energy when the temperature increases [244]. The ratio of valence electron to atoms
provides a good predictor of stacking fault energy, even when the alloying elements are changed.
Generally, the stacking fault energy shows a decreasing trend with the increase of the ratio of
valence electron to atoms [245].
In TWIP steels, manganese is the major alloying element. As an austenite stabiliser, a high
manganese content in the range of 15-30 wt.% makes TWIP steels fully austenitic at room
temperature [246]. The addition of alloying elements, such as carbon, silicon and/or aluminium,
is needed to obtain high strength and high uniform ductility associated with deformation-
induced twins in TWIP steels. Carbon stabilises FCC austenite phase and strengthens the steels.
The carbon contents in TWIP steels are either low, such as less than 0.05 wt.%, or high,
typically in the range of 0.5-1.0 wt.%, depending on the designed alloy system [246]. The
addition of approximately 0.6 wt.% carbon allows uniform, carbide-free, austenitic

44
microstructures and avoids the formation of martensite. The occurrence of TWIP effect is
related to the austenite phase stability. Austenite with a relatively low stability can progressively
transform to martensite during deformation. Enhancing the stability of austenite, then
suppressing the strain-induced austenite-to-martensite transformation during deformation of
TWIP steels, is essential in order to take full advantage of TWIP effect in these steels. Higher
carbon additions, however, may result in M3C carbide formation. Silicon improves strength by
solid solution strengthening. Silicon addition does not improve ductility, but it is effective in
refining martensite plates and increasing fracture strength [247]. By alloying with silicon, the
austenite-to-martensite transformation can be sustained. For some special purposes, e.g. to
improve the oxidation and corrosion resistance, chromium is also used in TWIP steels, along
with aluminium [248,249]. Vanadium is a strong carbide former. The addition of vanadium in
TWIP steels induces the formation of fine vanadium carbide precipitates, and increases the
threshold stress without affecting the hardening rate due to twinning [250]. The precipitates
have no significant effect on the desirable twinning effect in TWIP steels.
TWIP steels exhibit three hardening mechanisms during deformation, i.e. dislocation slip,
deformation twinning and martensitic transformation. A minimum strain is required for
generating twins in deformation [251]. As the strain exceeds the threshold at which the
deformation twins initiate, the amount of twins increases gradually with the progress of
deformation. Fig. 30 presents two electron backscatter diffraction maps performed on a Fe-22
wt.% Mn-0.6 wt.% C TWIP steel strained to 0.05 and 0.3 logarithmic strains [251]. It shows a
significant increase in the amount of deformation twins achieved by increasing the logarithmic
strain from 0.05 to 0.3. This provides more obstacles to dislocation slip and increases strength,
delays necking and fracture, and maintains a high work-hardening rate. For strains less than the
threshold, dislocation slip-slip interactions dominate the hardening. When strains exceed the
threshold, deformation twinning becomes the major mechanism, and slip-twin interactions are
responsible for hardening till the twinning volume fraction reaches a saturation level [252]. As
straining continues, transformation from austenite to martensite may take place if austenite is
not stable enough to suppress such phase transformation [253]. Fig. 31 shows the substructure
development in a 0.6 wt.% C-18 wt.% Mn-1.5 wt.% Al TWIP steel with the progress of
straining during tensile deformation [254]. It can be seen that only dislocations develop when
the strain is small, as shown in Fig. 31a. With the progress of straining, the substructure evolves
from a single variant twinning to two variants twinning, as shown respectively in Fig. 31b and c.
The mechanical twins formed in TWIP steels during plastic deformation act as barriers to
dislocation motion, and thus increase the strength of the steels.

45
(a) (b)

Fig. 30. Inverse pole figure electron backscatter diffraction maps of a Fe-22 wt.% Mn-0.6
wt.% C TWIP steel tensile deformed to (a) 0.05 and (b) 0.3 logarithmic strains [251].

(a) (b) (c)

Fig. 31. Bright field (a) and dark field (b,c) transmission electron microscopy images of grains
representing substructure development in a 0.6 wt.% C-18 wt.% Mn-1.5 wt.% Al TWIP steel
with strain of (a) 0.2, (b) 0.3 and (c) 0.4 [254]. GB: grain boundary.

The dominant deformation mode in TWIP steels is dislocation glide. However, as straining
continues the amount of deformation-induced twins increases—thus reducing the effective glide
distance of dislocations, which leads to the so-called dynamic Hall-Petch effect [246], as
schematically illustrated in Fig. 32. As the formation of mechanical twins involves the creation
of new crystal orientations, the twins progressively reduce the effective mean free path of
dislocations and increase the flow stress, resulting in a high strain-hardening behaviour. The
formation of deformation twins starts with nucleation. This requires a combination of critical
dislocation density coupled with maintaining relatively large homogeneous slip lengths [255].
Twin growth proceeds after the completion of nucleation by co-operative movement of
Shockley partials on subsequent {111} planes [251]. Possible mechanisms for the co-operative
movement include pole mechanism [256], cross-slip mechanism [257] and the reaction between

46
primary and secondary slip systems [258]. With the progress of deformation and the formation
of deformation twins, dislocation pile-up will occur at twin or grain boundaries, which increases
the strength of the steel.

Dislocation source Twin

Λ: dislocation mean free path

Fig. 32. Schematic illustration of the dynamic Hall-Petch effect [246].

TWIP steels combine extremely high strength with extremely high ductility. The mechanical
properties of typical TWIP steel grades are presented in Fig. 33 [246]. The strain-hardening
exponent (n) of TWIP steels increases to a high value of 0.4 at an approximate engineering
strain of 30%, and can then remain a constant afterwards, until total elongation reaches 50%.
The n-value of TWIP steels is much higher than that of other AHSSs. For example, the n-values
of TRIP 350/600 and DP 350/600 steels are respectively 0.2 and 0.15 [42]. The high rate of
strain-hardening associated with the deformation twinning phenomenon allows TWIP steels for
the combination of higher strengths and higher uniform elongations relative to other AHSS
grades, including TRIP and DP steels, as compared in Fig. 34 [259]. Details on the
microstructure-property relationships and the deformation behaviour of TWIP steels can be
found in two review articles of [260] and [261], respectively.
YS
Fe-18%Mn-C, Al UTS
Fe-25-31%Mn-Si, Al YS
Engineering stress (MPa)

UTS
Fe-22%Mn-C, N YS
UTS

Engineering strain (%)

Fig. 33. Mechanical properties of typical TWIP steels [246].


47
DP 980
TRIP 980 TWIP 980

Engineering stress (MPa)

TRIP 590
DP 590

Engineering strain (%)


Fig. 34. Engineering stress-strain curves of TWIP, TRIP and DP steels [259].

A major shortcoming of TWIP steels is their relatively low YS compared to other AHSSs,
which has limited their spreading use, particularly for automotive applications [262]. Solid
solution effect is commonly used to improve the YS of TWIP steels. It is recognised that carbon
is very effective for solid solution strengthening in TWIP steels, leading to increased YS
[263,264]. A number of studies have indicated that the addition of aluminium can induce
significant increase in YS of TWIP steels due to soloid solution strengthening. For example,
Jung et al. [265] found that the addition of aluminium increased the YS of TWIP steels by about
20 MPa per wt.% aluminium, and Jin and Lee [266] reported that the YS of TWIP steels could
be increased from 130 to 330 and 430 MPa when 1 and 2 wt.% aluminium was added,
respectively. In contrast, Hong et al. [267] found that the YS was slightly decreased from 449 to
440 MPa, even though 1 wt.% aluminium was added in TWIP steels, and an addition of 2 wt.%
showed almost no effect on the YS. Therefore, conflict of solid solution strengthening of
aluminium in TWIP steels still exists, and further investigations on the effects of aluminium on
the YS of TWIP steels are needed.
For a specific TWIP steel, four methods, including: (i) grain refinement; (ii) recovery and
partial recrystallisation of pre-strained steels; (iii) pre-straining by rolling; and (iv) precipitation
strengthening, can be adopted to enhance the YS [260]. Among these methods, pre-straining by
rolling may cause a severely reduced instantaneous strain-hardening coefficient and introduce a
significant anisotropy in the mechanical properties, which strongly reduce the formability of
TWIP steels. The high concentration of carbon, which is generally alloyed in TWIP steels, can
rapidly lead to the precipitation of iron carbides and, for longer annealing periods, the formation
of cementite. Precipitation of iron carbides induces decreased concentration of carbon in solid
solution, which lowers the stacking fault energy. Also, the cementite precipitated at grain
48
boundaries deteriorates the toughness and ductility of TWIP steels. Therefore, strategies for
avoiding iron carbide precipitation should be made in the processing of TWIP steels. Carbon
content plays a crucial role in the control of the appearance of iron carbides in TWIP steels.
Since carbon atoms tend to segregate at austenite grain boundaries [268], grain boundary
carbides can be precipitated easily during the cooling process. Cooman [269] suggested that the
maximum carbon content in high-manganese TWIP steels should be limited to around 0.6 wt.%
to avoid the precipitation of iron carbides. When the carbon content is increased to 0.7 wt.% and
above, the carbides precipitation kenetics will be remarkablely accelerated [260]. Aluminium
addition suppresses cementite precipitation due to the decrease in both activity and diffusivity of
carbon in austenite with aluminium addition [266,270]. The work of Hong et al. [271] showed
that the precipitation temperature of M3C carbides decreased from 700 to 635 ℃ when 1.9 wt.%
aluminium was added to an 18Mn TWIP steel, indicating that the addition of aluminium
suppressed the precipitation of M3C carbides. The strong attractive next-nearest-neighbour
carbon-aluminium interaction (25 kJ/mol) leads to an increase of the solubility limit of carbon
in austenite [261]. Since the nearest-neighbour carbon-aluminium interaction (-10 kJ/mol) is
repulsive, the formation of local order resembling the crystal structure of perovskite Fe 3AlC κ-
carbide will be favoured [272]. Manganese in TWIP steels hardly affects the precipitation of
M3C carbides [271]. In the presence of carbon, the nearest-neighbour silicon-carbon interaction
is strongly repulsive, which induces enhanced carbon activity in austenite. Bartlett et al. [273]
found that an increase in the amount of silicon from 0.59 to 1.56 wt.% in TWIP steels
accelerated the formation of κ-carbide, but did not increase the volume fraction. Silicon
increases the activity of carbon in austenite and stabilises the κ-carbide at high temperatures.
Therefore, methods (i) and (ii), which are commonly processed by a route combining cold
rolling and subsequent annealing, are widely used [274-277]. Accordingly, the majority of
previous research has focused on the cold-rolled and annealed TWIP steels [278,279]. During
cold rolling, mechanically induced twin boundaries facilitate the dynamic grain refinement.
Subsequent annealing leads to a reduced dislocation density, but preserves the fine-grained
microstructure due to the thermal stability of the deformed twins. However, the cold rolling and
annealing scheme has some shortcomings, including high energy consumption and complicated
processing. Consequently, TMP has become an effective alternative to refine and optimise the
microstructure through the control of the simultaneous dynamic restoration processes, including
DRX and DRV, directly in the processing of TWIP steels. TMP can improve the mechanical
properties of TWIP steels without changing steels’ chemical compositions and the extra step
required by annealing after cold rolling [280,281].
The hot working behaviour of TWIP steels is of primary importance for elaborating processing
methods consisting of hot rolling and subsequent cooling to room temperature. As the industrial

49
production of TWIP steels starts from hot rolling of the steels with a cast structure, Khosravifard
et al. [282] conducted TMP of two as-cast high-manganese TWIP steels with different levels of
carbon additions. They found that coarse austenite grains of the as-cast structure could be
effectively refined by DRX by applying optimised processing parameters. DRX generally initiates
at the prior austenite grain boundaries, in particular at the triple junctions with straining [280].
When higher strains are applied in TMP, deformation inhomogeneity, including distortion of grain
boundaries and twin boundaries, is increased, which promotes the nucleation of discontinuous
DRX. With the proceeding of deformation, mechanical twins are steadily formed within the initial
and DRX grains and at the grain boundaries, which play an important role in stimulating DRX
[283,284]. Mechanical twinning that occurs during TMP of a TWIP steel is affected by the steel’s
stacking fault energy and the deformation conditions. At higher strain rates and/or some other
loading conditions where high stress levels can be achieved, stacking fault energy may be
changed due to the effect of applied stress on the partial dislocation separation. In FCC materials
with low stacking fault energy, the applied stress changes the equilibrium separation distance of
the Shockley partials, and creates an effective stacking fault energy in austenitic steels [285,286].
The effective stacking fault energy can be quite low and may cause nucleation of mechanical
twinning, depending on texture and stress level [287,288]. Temperature has an impact on stacking
fault energy and is dependent upon various factors, including: solute or impurity pinning of
dislocations; Suzuki segregation to the stacking fault; and changes in short range order and local
or global chemical composition due to changes in the solubility of alloying elements with
variation of temperature [280]. Strain-induced precipitation or resolution may induce changes in
global or local compositions of TWIP steel, and thus contributes to stacking fault energy. It has
been reported that carbide precipitation at high temperatures depletes carbon from austenite matrix,
and leads to the decrease in stacking fault energy. A decrease in stacking fault energy facilitates
the mechanical twinning at high temperatures during TMP [289].
The microstructural evolution during recrystallisation is highly dictated by the characteristics
of deformed matrix, the orientation relationship between the deformed structure and growing
grains, and the temperature etc. By conducting double-hit TMP tests of a TWIP steel, Hajkazemi
et al. [290] showed that static recrystallisation during inter-pass time-induced softening, while
simultaneously, static strain-ageing as well as annealing twins during inter-pass time led to a
considerable hardening effect. An appropriate reduction applied in TMP is important to realise
DRX rather than DRV. The work of Grajcar and Borek [291] indicated that DRV would
dominate the strain-hardening process when a reduction of 20% was adopted. When the
reduction was increased to 40%, DRX became the main mechanism removing work hardening
of TWIP steels. Fine-grained austenitic microstructure of TWIP steels with a grain size up to
about 10 μm could be obtained through optimal design of TMP schedules by controlling the

50
metadynamic recrystallisation process [291].
Dobrzański et al. [292-295] conducted a lot of research on the multi-stage TMP of TWIP
steels with the purpose of understanding the effects of hot working conditions on
microstructural evolution. Fig. 35 shows a typical four-stage TMP schedule of a 27Mn-4Si-2Al-
Nb-Ti TWIP steel [292], which was designed to simulate the hot rolling process consisting of
several passes characterised by varying the amount of reduction and strain rate from pass to pass.
In their research, factors including temperature, strain and isothermal holding time after final-
stage deformation were found to have great effects on the behaviours of DRX and DRV during
TMP. A strain of 0.23 was sufficient to initiate DRX when deformation was conducted within a
high temperature range of 1050-1100 ℃, but the strain would be too low to initiate DRX at
950 ℃ or lower temperatures [292]. DRV would dominate the hot working process if the
deformation was performed at temperatures lower than 950 ℃ with a strain of 0.23, or the
whole temperature range if the smaller strain of 0.19 was applied to each of the four
deformation stages [293]. For continuous deformation, the initiation of DRX in the temperature
range from 850 to 1050 ℃ requires a true strain of at least 0.29 [294], and a true strain of about
0.51 is required to refine the microstructure through DRX [295]. The work of Dobrzański et al.
[292,293] has indicated that refined austenite grains with an average grain size of 10 μm can be
obtained through optimised TMP schedule when a true strain of 0.29 is applied to each of the
four deformational stages in combination of an isothermal holding time of 32 s after the final
deformation at 850 ℃.

Fig. 35. TMP schedule of a 27Mn-4Si-2Al-Nb-Ti TWIP steel [292].

Fig. 36 shows a TMP schedule of a Fe-Mn-(Al, Si) TWIP steel processed from 850 to 1100 ℃
through eight deformation stages with different strains and strain rates [296]. The designed
schedule can be used to simulate an eight-stand continuous tandem rolling process. The initial
austenitic structure of the TWIP steel to be processed had an average grain size of 50-60 μm. After
final deformation at 1100 ℃ and subsequent quenching in water, the steel was characterised by

51
uniform structure of DRX grains with size of 10-15 μm. Air cooling after final deformation did
not have significant effect on the microstructure, but a refined grain size of 7-10 μm could be
obtained when isothermal holding of 30 s was applied after final deformation as a result of
metadynamic and static recrystallisation [297-300], as indicated in Fig. 37.

Fig. 36. TMP schedule of a Fe-Mn-(Al, Si) TWIP steel with I: strain=0.4, strain rate=5 s-1; II
strain=0.3, strain rate=5 s-1; III strain=0.25, strain rate=10 s-1; IV strain=0.25, strain rate=10 s-1;
V strain=0.2, strain rate=50 s-1; VI strain=0.2, strain rate=50 s-1; VII strain=0.2, strain rate=100
s-1; and VIII strain=0.2, strain rate=100 s-1 [296]. TA-austenitising temperature, tiso-time of
isothermal holding of specimens at a temperature of the final deformation.

(a) (b)

(c) (d)

Fig. 37. Microstructure of Fe-Mn-(Al, Si) TWIP steel with (a) the initial microstructure of
austenite and (b,c,d) the microstructure of austenite after TMP shown in Fig. 36: (b) cooling
in water immediately after final deformation, (c) air cooling after final deformation, and (d)
cooling in water after 30 s isothermal holding [296].

TWIP steels with properly formed microstructure through optimally designed TMP schedules

52
provide the possibility of their applications in automotive structures after enhancement of their
strength levels. Refinement of austenitic grains improves not only the YS and UTS at room
temperature but also the cryogenic properties due to the suppression of the formation of ε- and α'-
martensite [301,302]. Khalesian et al. [303] conducted a research on the TMP of a high-
manganese TWIP steel and evaluated the room temperature mechanical properties of the
processed steel. In their study, TWIP steel was hot rolled respectively to total reductions of 30%
and 60% at temperatures of 800, 950 and 1100 ℃ with 10% reduction each pass through
controlled rolling. Their results indicated that the average grain size was refined as the rolling
temperature and equivalent strain were increased due to increased DRX driving force and
deformation twins formed during hot rolling. Twinning can provide additional nucleation sites
for DRX, therefore promoting the DRX process [280,290,304]. Fig. 38 shows the room
temperature tensile properties of a high-manganese TWIP steel after TMP-controlled hot rolling
with different reductions [303], which indicates an increase in strength after TMP as compared to
the as-received material without a dramatic decrease in elongation to facture values. Mechanisms
including non-recrystallisation, partial recrystallisation and complete recrystallisation dominate
the evolution of microstructure and the change of gain size, which have contributed to the
alteration of mechanical properties of the TWIP steel during TMP [305-307].
Engineering stress (MPa)

Engineering stress (MPa)

Elongation (MPa)
Elongation (MPa)

As-received 800 ℃ 950 ℃ 1100 ℃ As-received 800 ℃ 950 ℃ 1100 ℃


(a) (b)
Fig. 38. Room temperature tensile properties of a high-manganese TWIP steel after TMP-controlled
hot rolling with reductions of (a) 30% and (b) 60% [303]. UE: uniform elongation.

Recent research conducted by Belyakov et al. [308] indicated that an average grain size of 26
μm could be obtained when high-manganese TWIP steel was hot rolled at 1100 ℃ due to the
mechanism of discontinuous post-DRX [309]. Warm rolling at 500 ℃ could result in a much finer
grain size of 10 μm in the transverse direction with the deformation grains being flattened and
highly elongated along the rolling direction. A decrease in rolling temperature from 1100 to
500 ℃ resulted in significant increase in YS from 362 to 950 MPa, while UTS did not increase
substantially (from about 1100 to 1310 MPa). As the elongation of the warm-rolled steel
53
maintained a high level (34%), rolling under warm working conditions could be a promising TMP
method for production of TWIP steels with beneficial combination of mechanical properties [308].
A similar finding was obtained by Lee et al. [310] on the calibre rolling of a Fe-17Mn-0.6C TWIP
steel through controlled TMP at 500 ℃. The results showed that uniform elongation of the warm-
rolled TWIP steel was not severely degraded, although its strength had been greatly enhanced
through grain refinement in the transverse direction due to the occurrence of deformation twinning,
dynamic strain ageing and formation of stacking faults [311-314].
The TMP of TWIP steels is of practical significance for the implementation of industrial
production. However, the low plasticity and high deformation resistance during TMP-controlled
hot or warm rolling of TWIP steels due to high contents of manganese, silicon and aluminium
may cause some difficulties. For example, severe edge cracks may be formed during hot strip
rolling, leading to the edge cutting-off of the final products [315]. Also, the thickness of the
steel sheet cannot be controlled exactly and the microstructure cannot stay homogeneous
through hot or warm rolling when compared to cold rolling. This may influence the final
mechanical properties of the products [310,316]. Nevertheless, TWIP steel products with high
surface quality and dimensional accuracy, as well as excellent combinations of mechanical
properties, could be obtained through proper control of the processing conditions based on the
optimised design of TMP schedules.

3.2 Thermomechancal processing of AUST.SSs


AUST.SS is regarded as one kind of AHSS because this steel grade provides not only
exceptional corrosion resistance but also excellent mechanical properties when compared to
conventional carbon steels. As the name suggests, AUST.SS has an austenitic FCC crystal
structure, which is very tough and ductile, even at sub-zero temperatures. AUST.SS is a solid-
solution alloy and each alloying element plays an important role in the control of austenite
stabilisation, corrosion resistance and mechanical properties. It contains chromium, manganese,
silicon and carbon. It may also contain nickel, molybdenum, titanium, niobium, copper and
some other minor alloying elements. Chromium is an essential element in stainless steels,
increasing oxidation resistance and corrosion resistance. In AUST.SS, its content is generally
between 16 and 26 wt.% [4]. Manganese and carbon are both nickel equivalents, which can
increase austenite stability when added to AUST.SS in appropriate matching amounts. Nickel is
an austenite stabiliser and can significantly improve resistance to oxidation and corrosion when
added to AUST.SS. A number of traditional AUST.SS grades in the 2xx and 3xx series contain
a nickel content of 4.5-12.1 wt.% [317]. For example, the AUST.SS grade 304—the most
familiar and most widely used alloy in the austenitic group—contains 8-10 wt.% nickel [4].

54
However, due to the high cost of nickel and price fluctuations, in recent years reduced nickel-
containing and nickel-free AUST.SSs have been developed as an alternative to the expensive
austenitic chromium-nickel series [318,319]. When the amount of nickel is reduced or totally
excluded, nitrogen, manganese, copper and carbon can be used in lieu of nickel to maintain an
austenitic microstructure with an acceptable balance between the production cost and properties
of AUST.SS [318,320]. Nitrogen is a strong austenite stabiliser. It has a beneficial effect on
mechanical properties and strain-hardening behaviour as long as the solubility limit of nitrogen
in austenite is not exceeded. It has been found that high-nitrogen AUST.SS can achieve the
excellent combination of mechanical properties and corrosion resistance at a relatively low
production cost, by replacing nickel in conventional Fe-Cr-Ni AUST.SS with nitrogen. The
reduction of nickel is also beneficial in suppressing environmental contamination [321].
Molybdenum is commonly used in AUST.SS to improve resistance to chloride pitting
through formation of compact passive films on the steel surface [322]. It reduces the intensity of
the oxidising effect required to insure passivity and decreases the tendency of previously formed
passive films to break down. Molybdenum also increases hardenability, slows the critical
quenching speed, improves the high temperature strength and hardness, and enhances the creep
strength of AUST.SS at elevated temperatures. Molybdenum is a ferrite stabiliser which, when
used in AUST.SS, must be balanced with austenite stabilisers in order to maintain the austenitic
microstructure. Silicon in AUST.SS promotes oxidation resistance and resistance to corrosion
by oxidising acids. In small amounts, silicon confers mild hardenability on steels. Small
amounts of silicon and copper are usually added to AUST.SS containing molybdenum to
improve corrosion resistance in sulphuric acid. Titanium and niobium are used as stabilising
elements in AUST.SS to prevent sensitisation to intergranular corrosion following a sojourn of
the steel within the temperature range in which precipitation of chromium carbide may occur.
The theoretical amounts of titanium and niobium required for a full stabilisation based on
stoichiometric calculations can be described by Eqs. (2) and (3), respectively [323]:

Ti ≥ 0.15 + 4 (%C + %N) (2)


Nb ≥ 0.2 + 5 (%C + %N) (3)

The microstructural characteristics of AUST.SS are dependent upon the chemical


compositions, processing history and heat treatment technologies, such as annealing, being
applied after processing. Fig. 39 shows a scanning electron microscope image of austenite
grains of an annealed 304 type AUST.SS [324]. It shows an austenitic microstructure contains a
number of twins inside grains. These are caused by a certain type of shearing that develops

55
inside the grain during plastic deformation. Unlike ferritic steels, which have dramatic
microstructural changes that depend on the peak operational or failure temperature, there are no
abrupt microstructural changes in the AUST.SSs [324]. Fig. 40 shows the electron backscatter
diffraction map of a nickel-free AUST.SS after multi-pass hot rolling at 900 ℃, in which a fine
austenite grain size of about 13 µm is observed [325]. The refinement of austenite grains
indicates that DRX has occurred in some regions as a result of high reduction during hot rolling.

Fig. 39. Scanning electron microscope image of a 304 type AUST.SS strain-annealed at
927 ℃ for 72 h [324].

Fig. 40. Electron backscatter diffraction map of a Ni-free AUST.SS after multi-pass hot rolling
at 900 ℃ [325].

Hot and cold rolling can be used to produce thin sheet AUST.SS products through a
reduction in thickness of cast or hot rolled slabs. In order to eliminate the work-hardening effect
caused by rolling, annealing is usually applied to control the grain size, modify microstructure
and dissolve precipitated carbides and/or nitrides. The mechanical properties of AUST.SSs can
be improved by careful design of annealing treatment procedures. The work of Hamada et al.
[318] showed that controlled reversion annealing technology could be applied to generate
ultrafine austenite grains of heavily cold-worked AUST.SS. A combination of grain refinement,

56
and TRIP and TWIP effects can significantly contribute to the improvement of mechanical
properties of AUST.SSs to make them superior to the conventional annealed or work-hardened
stainless steels, as shown in Fig. 41 [318].

Total elongation (%)

Tensile strength (MPa)


Fig. 41. Comparison of the mechanical properties between the reversion annealed type
201 AUST.SS and other conventional steel grades [318].

AUST.SSs cannot be hardened by heat treatment, but can be strengthened by work hardening
to obtain high YS and UTS, while retaining good ductility and toughness, even at cryogenic
temperatures. They do not exhibit a clear yield point and their mechanical behaviour is
dependent upon the strain rate. Fig. 42 shows tensile curves of a type 304 AUST.SS at room
temperature obtained under different strain rates [326]. It is clear that the YS is increased from
300 to 340 MPa, whereas the UTS is decreased from 900 to 850 MPa when the strain rate is
increased from 3 × 10-3 to 3 × 10-1 s-1. This change in tensile strength is thought to be caused by
the increased amount of non-planar slips induced by the heat from deformation when the strain
rate is increased, thereby inhibiting transformation due to the reduction of nucleation sites [326].
AUST.SS grades are not subject to an impact transition at low temperatures and possess high
toughness to cryogenic temperatures due to the multiplicity of slip system in the FCC austenite.
The excellent toughness allows AUST.SS to be an exceptional cryogenic material that can be
used in various low temperature environments, especially for energy-absorbing components.
Tensile strength (MPa)

True strain (%)


Fig. 42. Tensile curves of a type 304 AUST.SS obtained under different strain rates [326].
57
AUST.SSs are among the most widely used structural materials because of their beneficial
combination of mechanical properties and corrosion resistance. However, a common
disadvantage of AUST.SSs is their relatively low YS levels, which limits their wide usage in
structural applications and the automotive industry [327,328]. Two methods are generally used
to improve the YS of an AUST.SS: adjusting chemical compositions to induce hardening by
both substitutional and interstitial solid solutions; and refining the austenite grain size [329]. For
a given AUST.SS with specified chemical compositions, grain refinement is the most efficient
way to achieve the satisfied mechanical properties. It is known that AUST.SS exhibits a single-
phase FCC structure that is maintained over a wide range of temperatures. Unlike steels (e.g.
ferritic steels), which can be refined by phase transformations, the grain refinement in AUST.SS
is primarily achieved by recrystallisation in hot working or heat treatment through the control of
processing parameters due to the absence of phase transformation. Application of TMP on
AUST.SS is of practical significance because AUST.SS products are commonly fabricated by
various TMP-controlled processes such as rolling, forging and extrusion. TMP can offer
AUST.SS grain refinement and substructure hardening that increase strength without
significantly impairing ductility and toughness.
AUST.SSs are characterised by their low-to-medium stacking fault energy which facilitates
the accumulation of moving dislocations. The key restoration phenomenon during and after hot
deformation of AUST.SSs is DRX [330-332]. This is technologically important as it reduces
deformation resistance and can bring about significant grain refinement, which in turn improves
the mechanical properties and formability of the steels [333,334]. DRX takes place at elevated
temperatures after a critical strain is attained [335,336]. Up to the critical strain, dislocation
density and defects increase as the deformation level increases. Once the critical strain is
achieved, further deformation of the material leads to strain-free grains nucleating
predominantly at grain boundaries. More strain-free grains nucleate as straining progresses and
they grow until the entire deformed structure is replaced by newly recrystallised grains [337].
The critical strain for DRX initiation depends on the chemical compositions of AUST.SS,
the size of austenite grains, and the TMP schedules (deformation temperature and strain rate).
Mandal et al. [338] studied the DRX behaviour of an as-cast AUST.SS 304 by using TMP
technology, and found that the DRX process of coarse austenite grains was relatively slow when
compared to fine austenite grains. The boundaries of fine grains can offer DRX more nucleation
sites. In fact, in the microstructure with finer grain size, DRX can start faster and proceed with
higher kinetic energy [339]. Moreover, precipitates formed during the TMP of AUST.SSs may
also affect the progress of DRX. The work of Sun et al. [340] showed that coarsened
precipitates at high angle grain boundaries facilitated the nucleation of DRX grains, but

58
simultaneously prohibited the growth of DRX grains. Mandal et al. [338] indicated that
austenite could be refined completely by DRX under high deformation temperature (1100-
1200 ℃) and slow strain rate (0.5-5 s-1) conditions. This is confirmed by Han et al. [305], who
found that high temperature and low strain rate facilitate the formation of complete DRX. To
accelerate the progress of DRX, the deformation temperature should be increased; and to
significantly refine grain size, a high strain rate should be adopted [341]. While it should be
noted that an increased strain rate can reduce the size of recrystallised grains, it may also induce
incomplete DRX due to insufficient time for DRX formation during the high strain rate
deformation. Therefore, an optimal strain rate should be selected in the TMP of AUST.SSs
based on the requirement of size and volume faction of DRX grains.
In order to evaluate the critical condition for onset of DRX, Samantaray et al. [342] studied
the flow behaviour and microstructural evolution of a 316 L(N) AUST.SS during TMP. They
pointed out that DRX would not be initiated until a critical true plastic strain of 0.45 was
reached at the temperature of 1000 ℃ and strain rate of 1 s-1, and necklace DRX grains
nucleated only in the specimen deformed to an engineering strain of 50% at 1000 ℃/1 s-1. For
other processing conditions, e.g. a strain rate less than 1 s-1, temperature lower than 1000 ℃ and
engineering strain less than 50%, DRX would not be witnessed or completely formed. This
finding is consistent with the results obtained by other researchers on the TMP of diverse
AUST.SS grades, i.e. DRX occurs more readily at high temperature and high strain rate
[305,339,340]. The critical strain is not a constant but varies according to the deformation
temperature and strain rate. A number of researchers have indicated that the critical strain for
initiating DRX in AUST.SSs is likely to increase when the deformation temperature decreases
and the strain rate increases. [341,343].
Depending on deformation conditions, AUST.SSs may exhibit different DRX behaviours.
Yanushkevich et al. [344] conducted TMP-controlled rolling tests on a 316L-type AUST.SS
within a temperature range of 500-900 ℃, and found that discontinuous DRX dominated the
microstructural evolution at 900 ℃. At 500 ℃, the main mechanism was continuous DRX with
a well-developed spatial sub-boundary net indicated. Some portions exhibited high-angle
misorientations and looked like incomplete grain boundaries in the grain interiors.
Discontinuous DRX takes place in a cyclic manner under hot working conditions, when new
grains repeatedly nucleate due to local migration, i.e. grain boundaries bulge, then
recrystallising nuclei grow at the expense of work-hardened surroundings. Differently,
continuous DRX occurs under warm working conditions. In this case, the new grains evolve as
a result of progressive increases in misorienations among the strain-induced sub-boundaries up
to typical values of high-angle grain boundaries [309]. Compared to discontinuous DRX under

59
hot working, continuous DRX can be used to achieve significant grain refinement through the
application of large strain deformation with high flow stress levels under warm working
conditions [345,346]. A comparative study showed that the transverse grain size decreased from
3.4 to 0.85 µm as the rolling temperature decreased from 900 to 500 ℃, which brought about an
increase of YS from 720 to 945 MPa at room temperature [344].
As AUST.SSs for structural and automotive applications are primarily produced by rolling, a
number of research studies have been conducted on TMP-controlled rolling processes. Ikeda et
al. [347] studied the effect of the TMP schedule on strengthening of a 22Mn-13Cr-5Ni
AUST.SS for cryogenic use. They found that the finish-rolling temperature, cooling rate and
start-cooling temperature were important factors controlling the size of recrystallised austenite
grains. The grains became finer as the finish-rolling temperature decreased, and the cooling rate
and start-cooling temperature increased in the recrystallisation region. In this case, a decrease in
the finish-rolling temperature plays a role in refining the recrystallised austenite grains. An
increase in the cooling rate and start-cooling temperature helps to restrain the growth of
recrystallised austenite grains. The results of a study by Ikeda et al. [347] indicated that the
austenite grain size of AUST.SS can be refined up to 10 μm by applying TMP, which
contributed to YS increases of 50 MPa at room temperature and 100 MPa at -269 ℃ when
compared to the solution heat-treated steel. Han et al. [348] conducted a systematic
investigation on the effects of rolling parameters, including finish-rolling temperature, cooling
rate and coiling temperature, on the microstructural evolution and mechanical properties of a
304 AUST.SS by controlled TMP, as schematically shown in Fig. 43. A rising of finish-rolling
temperature from 1000 to 1050 ℃ was found to induce increased fraction of recrystallised
grains with improved homogeneous distribution, and an increase in cooling rate from 10 to
30 ℃/s would lead to much finer grains. Therefore, finish-rolling temperature and cooling rate
are two crucial parameters controlling the grain size and grain homogenisation. Also, the coiling
temperature after hot rolling was found to affect the grain size: finer average grain size could be
attained when coiling at the lower temperature [348].

60
Start-rolling temperature: 1200 ℃
Finish-rolling temperature: 1000-1050 ℃
Strain rate: 10 s-1
1250 ℃, 5 min Total reduction: 60%
Temperature (℃) 5 ℃/s
5 ℃/s
60 s
60 s Cooling rate: 5 ℃/s
10 ℃/s
650 ℃, 60 s
Air cooling

Time (s)
Fig. 43. Schematic diagram of TMP-controlled hot rolled process of a 304
AUST.SS [348].

The results obtained by Han et al. [348] indicated that a refinement of grain size by DRX not
only improved the tensile strength—particularly YS—but also increased the intergranular
corrosion resistance of AUST.SS through application of appropriate TMP technology.
Intergranular corrosion is caused by the precipitation of carbides at grain boundaries, and the
corrosion rate is affected by the volume fraction of precipitated carbides per unit of grain
boundary area. As grain size is refined, the grain boundary area’s per unit volume increases and
the degree of the chromium depletion caused by carbide precipitation decreases for a given
carbon content. Hence, boundaries may not be sensitised in finely grained materials, which can
therefore increase the intergranular corrosion resistance [349,350]. The work of Penttilä et al.
[351] on an AUST.SS 316L showed that grain refinement induced an increase in oxidation
resistance due to the formation of a very thin chromium-rich oxide film on the steel’s surface.
Abe et al. [352] investigated the effect of grain refinement on the oxidation behaviour of
austenitic stainless steels in superheated steam. They reported an improvement in oxidation
resistance due to a fine-grained layer forming on alloy surfaces. They believed the fine-grained
structure was responsible for forming the protective chromium oxide (Cr 2O3) layer. Nezakat et
al. [353] applied TMP technology to improve oxidation resistance by refining the microstructure
of an AUST.SS 316L. Their results showed that rolling reduction should be 50% or more to
totally prevent oxide exfoliation. They successfully refined the austenite grain size from 20 to 1
μm via TMP with almost the same texture, and indicated that the oxidation resistance could be
enhanced by grain size refinement. There is a critical grain size for the formation of a protective
chromium oxide layer. This layer decreases the diffusion rate of iron ions and oxygen ions—and,
as a result, reduces the oxidation rate—when the fraction of grain boundaries is relatively high.

61
The critical grain size varies between AUST.SSs. For example, it is 3 μm for an AUST.SS 316L,
but 8 μm for an AUST.SS alloyed with copper [354]. Therefore, different TMP schedules
should be applied to AUST.SSs with distinct chemical compositions in order to improve their
oxidation resistance through grain refinement.
Hot workability is a major concern for industries because hot deformation is the primary
processing route for the production of various components of AUST.SSs. In TMP, some
processing regimes may not be favourable for hot deformation—even though they are effective
for grain refinement by DRX. This is due to instabilities, such as localised deformation,
adiabatic shear band formation, cavitation, hot shortness or wedge cracking, which can lead to
deterioration of the mechanical properties of the final product [355]. Hot workability of
AUST.SSs is therefore concerned with plastic deformation having a defect-free microstructure
during a specific TMP process. A safe window of optimal processing parameters is required not
only to produce a fine and equiaxed DRX microstructure, but also to avoid defects and flow
instability within it.
Processing maps based on the dynamic materials model are widely used to study the hot
workability of AUST.SSs during hot deformation. They can evaluate the deformation
mechanisms of determined regions, and also describe the instability regions that should be
avoided during hot deformation process [343,356,357]. Sun et al. [358] conducted TMP to
investigate the optimal hot deformation parameters of a modified 310 AUST.SS using
processing maps, and found that deformation in the safe region was beneficial to DRX and
DRV, while deformation in unstable region would lead to flow instability, kink boundaries and
grain growth. Cheng et al. [359] conducted a double-hit TMP on a HR3C heat-resistant
AUST.SS over a temperature range of 950-1150 ℃, as schematically shown in Fig. 44a. They
found that there was one stable domain with peak power dissipation efficiency in the region of a
deformation temperature of 1150 ℃ and strain rates of 0.1-1 s 1 at the strain of 0.6, as shown in

Fig. 44b. Han et al. [305] found that DRX of 904L AUST.SS could occur sufficiently within the
temperature and strain rate ranges of 1050-1125 ℃ and 0.05-1 s-1 or of 1130-1150 ℃ and below
0.05-1 s-1, respectively, where the peak power dissipation efficiency, along with stable flow,
could be achieved. If deformation was performed in the regions of low temperature and high
strain rate, flow instability characterised by severe shear bands with localised DRX grains
would occur, as shown in Fig. 45, will be caused.

62
(a) (b)

Fig. 44. (a) Schematic diagram of double-hit TMP, and (b) processing map at strain of 0.6
indicating the unstable (grey area) and stable (dashed rectangle) flow domains of a HR3C
heat-resistant AUST.SS [359].

Fig. 45. Flow instability occurred in a 904L AUST.SS at deformation


temperature of 1050 ℃ and strain rate of 10 s-1 [305].

Generally, an increase in power dissipation efficiency indicates an enhancement in flow


stability, i.e. the higher the power dissipation efficiency, the more stable the flow during hot
deformation [360]. In AUST.SSs, DRX is often regarded as the clear sign of stable flow, and
the grain refinement that occurs during DRX is responsible for the increment in power
dissipation efficiency [356,361-363]. However, in some cases a high power dissipation
efficiency may also induce unstable flow, depending on the steel’s type and the processing
conditions. Sahu et al. [364] established processing maps of a Fe-23wt.%Cr-8wt.%Ni AUST.SS
using TMP technology in the temperature range of 950-1100 ℃ and strain rate rage of 0.01-10 s-
1
. They found that the peak efficiency of 35.91% at strain of 0.6 could be obtained under a
temperature of 1100 ℃ and strain rate of 1 s-1. This is in agreement with the power dissipation
efficiency (max. 30%-40%) for DRX in AUST.SSs [305,365,366], which indicates that the
highest amount of DRX should have taken place in the peak efficiency domain. Higher
efficiency corresponds to the greater energy involved in microstructural change, and results in

63
DRX grains that are beneficial to the hot deformation process. The optimum processing
condition can be identified as the value of highest efficiency, but it should be in the stable or
safe processing zone. However, as the processing parameters corresponding to peak efficiency
are located outside the stable region of the processing map, as shown in Fig. 46 [364], the
domain with the peak efficiency of 35.91% should be avoided. Twin formation and necklace
microstructure are thought to be the dominating damage mechanisms that have caused such
microstructural instability [364]. Therefore, an optimal design of TMP schedules should be
determined only with a comprehensive investigation of both the microstructural evolution and
the flow behaviour with the purpose of taking full advantage of DRX in the grain refinement of
AUST.SSs.

Fig. 46. Processing map of a Fe-23wt.%Cr-8wt.%Ni AUST.SS at strain of 0.6 [364]. The
shaded regions are the stable domains.

These TMP routes can achieve austenite grain refinement through appropriate control of the
DRX process. However, the level of refinement is limited due to the relatively high DRX
temperature. Grain size is usually in the domain of several or tens of micrometres. As a result,
the mechanical properties of AUST.SSs cannot be significantly improved by conventional TMP.
Limited improvement of the mechanical properties of AUST.SSs—especially the combination
of YS and ductility—imposed by conventional TMP can be surpassed by an alternative TMP
route of martensite-to-austenite reversion. In this approach, metastable austenite is first severely
cold rolled at room temperature to induce the formation of martensite. Subsequent annealing is
applied to realise grain refinement through martensite-to-austenite transformation. Severe
deformation of metastable austenite leads to strain-induced austenite-to-martensite
transformation. On annealing, the severely deformed strain-induced martensite reverts back to
austenite.
Reduction in cold rolling and the subsequent annealing conditions have significant effects on

64
the martensite-to-austenite reversion process, and thus affect the level of austenite grain
refinement. Metastable austenite should be almost completely transformed to martensite during
cold rolling at room temperature. This is because the remaining untransformed austenite would
not contribute significantly to the refinement of austenite grains after martensite-to-austenite
reversion [367]. For a specific AUST.SS, the amount of martensite induced by cold rolling is
dependent on the reduction. The work of Tomimura et al. [367] on Fe-Cr-Ni AUST.SSs showed
that more than 90% of austenite could be transformed to martensite by 90% cold rolling when
the nickel equivalent in steels was less than 16 wt.%. Naghizadeh et al. [368] found that the
amount of achievable strain-induced martensite in a 304 AUST.SS was saturated when 70%
reduction was applied in cold rolling. In the work by Somani et al. [369], cold rolling with 55%
reduction was found to be adequate to realise nearly 100% martensite in 301LN AUST.SS.
While for the 304L AUST.SS, a reduction of 50% was sufficient to induce almost completely
transformed martensite during cold rolling [370]. Therefore, the cold rolling reductions required
for inducing completely transformed martensite vary according to the AUST.SSs. The amount
of strain-induced martensite from metastable austenite during cold rolling can be estimated
using the Olson-Cohen equation [371]:

f    1  exp{ [1  exp( )]n} (4)

where f   is the volume fraction of martensite, α and β are are temperature-dependent


parameters, ε is the plastic strain, and n' is a fixed exponent 4.5. The α parameter defines the
course of shear-band formation with strain, and is temperature sensitive through its dependence
on stacking fault energy. The β parameter is proportional to the probability than an intersection
will form an embryo, and this probability is temperature dependent through its relationship to
the chemical driving force.
In order to maximise the austenite grain refinement, the type of martensite structure should
be adjusted. When the reduction in cold rolling is low, lath-type martensite dominates the
microstructure. A further increase in cold rolling reduction leads to martensite fragmentation,
and a change in the predominant morphology of martensite from lath-type to dislocation-cell-
type. A schematic diagram of the transition from lath-type martensite to dislocation-cell-type
martensite during severe cold rolling is illustrated in Fig. 47 [372]. If the martensite structure is
lath-type, austenite nucleates at the lath boundaries and intersections, and grows from these sites
into the martensite lath during martensite-to-austenite reversion treatment. If the martensite
morphology is changed from lath-type to dislocation-cell-type, the defect density would
increase considerably, and equiaxed ultrafine austenite grains would nucleate at martensite grain
65
boundaries. Misra et al. [372] found that both lath-type and dislocation-cell-type martensite
structures were formed in a 301LN AUST.SS when a cold-rolling reduction of 45% was applied,
but a further increase in the reduction to 77% led to predominantly dislocation-cell-type
martensite structure. As a result, the refined austenite grain size in the reduction of 45%-rolled
and 77%-rolled samples was 1.1 and 0.6 μm, respectively, after martensite-to-austenite
reversion under the same annealing condition. Thus, to achieve a high level of grain refinement,
the dislocation-cell-type martensite is ideal because it contains a much higher number of
nucleation sites for the austenite grains during martensite-to-austenite reversion annealing, when
compared to the lath-type martensite.

Fig. 47. Schematic illustration of the transition from lath-type martensite to dislocation-
cell-type martensite during severe cold rolling [372].

The kinetics of martensite-to-austenite reversion is dependent upon both the degree of prior
cold rolling reduction and the annealing conditions. A higher reduction in cold rolling induces
the reversion process faster. Somani et al. [369] found that the martensite-to-austenite reversion
in the reduction of 75%-rolled 301LN AUST.SS could be nearly completed when annealing at
700 ℃ for only 6 s. By contrast, 10% of martensite remained in the reduction of the 45%-rolled
sample, even when it was held at 700 ℃ for 100 s. The annealing temperature plays a
substantial role in the martensite-to-austenite reversion process. An increase in the annealing
temperature enhances the rate of martensite reversion and the degree of coarsening of reverted
austenite grains [373]. Austenite grain size after martensite-to-austenite reversion increases with
an increase in the annealing temperature. This is due to significant grain growth at high
annealing temperatures and it is difficult to control the reversion annealing to achieve much
finer austenite grains [368]. Moreover, the formed austenite after martensite-to-austenite
reversion is partly transformed back to martensite at high annealing temperatures, thus
deteriorating the final mechanical properties of the processed AUST.SSs [374].
If the aim is to decrease the amount of martensite during annealing, then high temperatures
should be avoided. The increased volume fraction of martensite at high temperatures is due to

66
the precipitation of carbides at grain boundaries. This process depletes carbon from the austenite
matrix and leads to an increase in the Ms temperature, and thus causes thermally induced
martensite formation during cooling [374-377]. The holding time in annealing affects the
martensite-to-austenite reversion process and the size of reverted austenite grains. Generally, a
higher annealing temperature requires less holding time to achieve the complete reversion, and
an increase in the holding time leads to increased austenite grain size after martensite-to-
austenite reversion [368,370,373].
The reversion of martensite in AUST.SSs takes place at a temperature much lower than the
recrystallisation temperature of the steels. The reversion mechanism significantly depends on
the chemical compositions of steels and the annealing temperature. Depending on the
temperature, the strain-induced martensite can be metastable and its reversion to austenite may
occur by the diffusional or the shear mechanism. The diffusional reversion mechanism is
characterised by: (i) a wide annealing temperature range where the reversion occurs; (ii) the
formation of defect-free equiaxed austenitic grains which grow in size with time; (iii) a wide
austenite grain size distribution; and (iv) possible formation of the secondary phase precipitates
[378]. In a shear reversion mechanism, the reversion occurs within a narrow range of
temperatures, and becomes a diffusional-dependent process only after the initial shear reversion
of martensite to high-dislocation-density austenite occurs [374]. The shear reversion process
first involves transformation of strain-induced martensite to reverted austenite laths with a high
density of defects. With an increase in annealing temperature or holding time, the defect-free
austenite sub-grains are formed. In the later stage, sub-grains coalesce to form a
nanograined/ultrafine-grained structure resembling that of a recrystallised structure [379]. The
shear-reverted austenite contains a high density of dislocations immediately after the reversion
and the austenite grains are refined through recovery and recrystallisation, while the diffusional-
reverted austenite is characterised by the nucleation of equiaxed grains within the martensite
matrix and the austenite grains gradually grow during annealing [367]. In practice, the
diffusional mechanism is dominant for the martensite-to-austenite reversion at relatively low
annealing temperatures, and the shear mechanism is generally operative when the annealing
temperature is high [373].
Based on the above analysis, the following four conditions should be satisfied in order to
effectively refine the austenitic structure in AUST.SSs by means of martensite-to-austenite
reversion treatment: (i) metastable austenite should be completely transformed to martensite by
cold rolling; (ii) the cold rolling reduction should be high enough to induce transition from lath-
type martensite to dislocation-cell-type martensite; (iii) the annealing temperature for
martensite-to-austenite reversion should be as low as possible to suppress coarsening of the

67
reverted austenite grains and transformation back of the reverted austenite to martensite; and (iv)
a short period of annealing time should be used to inhibit grain growth of the reverted austenite.
The TMP schedule of AUST.SSs involving martensite-to-austenite reversion is schematically
illustrated in Fig. 48. Ultrafine austenite grains can be achieved through careful design of the
martensite-to-austenite reversion TMP schedules and optimisation of the processing parameters.
A large number of researchers have used this approach to successfully refine the austenite grains,
and thus improve the mechanical properties, especially YS and elongation, of AUST.SSs
through the control of the processing parameters, as summarised in Table 1.

Dislocation-cell-type martensite Annealing


Metastable austenite Lath-type martensite Rs Reversion
Rf

Refined austenite grains


Cold rolling Cold rolling
Fig. 48. Schematic illustration of the TMP schedule of AUST.SSs involving martensite-to-austenite
reversion. Rs and Rf are the reversion starting and finishing temperatures, respectively.

Table 1 Austenite grain size and mechanical properties of typical AUSS.SSs after
martensite-to-austenite reversion TMP treatment.
AUST.SSs Cold rolling Annealing Holding time Grain size YS Elongation References
reduction temperature
316LN 90% 900 ℃ 120 s 2 μm 994 MPa 40% [380]
201 60% 800 ℃ 10 s 1.5 μm 800 MPa 50% [318]
15Cr-9Mn 60% 800 ℃ 10 s 1.2 μm 542 MPa 63% [381]
301LN 62% 800 ℃ 1s 0.7 μm 711 MPa 37% [369]
304L 65% 850 ℃ 60 s 0.62 μm 855 MPa 44% [373]
301LN 77% 800 ℃ 1s 0.6 μm 880 MPa 44% [372]
15Cr-9Mn-0.11Nb 60% 800 ℃ 10 s 0.6 μm 813 MPa 51% [381]
301LN 63% 800 ℃ 1s 0.54 μm 700 MPa 35% [382]
16Cr-10Ni 90% 600 ℃ 600 s 0.5 μm 700 MPa - [367]
304L 90% 700 ℃ 18,000 s 0.33 μm 1000 MPa 40% [370]
16Cr-10Ni 74% 850 ℃ 100 s 0.25 μm 585 MPa 35% [383]
18Cr-12Mn-0.25N 80% 900 ℃ 100 s 0.24 μm 1150 MPa 21% [384]
201Nb 90% 900 ℃ 60 s 0.093 μm 1000 MPa 35% [385]
304 85% 580 ℃ 1,800 s 0.08 μm 1120 MPa 12% [386]
201L 95% 850 ℃ 30 s 0.065 μm 1300 MPa 33% [387]

3.3 Thermomechancal processing of L-IP steels


The idea of developing L-IP steels comes from the increasing demand for vehicle weight
reduction with the purpose of reducing greenhouse gas emission and improving fuel efficiency
[388,389]. L-IP steels belong to the Fe-Mn-Al-C system, which contains high levels of
manganese, aluminium and carbon, in which aluminium is the key alloying element to realise
the “lightweight” concept. L-IP steels can be basically divided into three categories, namely:
68
austenite-based; ferrite-based; and “austenite + ferrite” duplex steels, depending on their matrix
phase constituents. L-IP steels differ from TWIP steels in microstructure because TWIP steels
contain only a single austenite phase due to the very high content of manganese, even though 2-
3 wt.% aluminium is added [390,391]. Manganese and carbon are both austenite formers, and
aluminium stabilises ferrite and increases the metastable solubility of carbon by lowering the
diffusivity. Because carbon promotes carbide formation, a reasonable design of chemical
compositions for L-IP steels can generate triplex phases that are composed of austenite, ferrite
and κ-carbide (Fe,Mn)3AlC [392]. A scanning electron microscope image of an “Fe-12 wt.%
Mn-5.5 wt.% Al-0.7 wt.% C” L-IP steel is shown in Fig. 49, from which the microstructure of
austenite, ferrite and κ-carbide can be clearly observed.

Fig. 49. Scanning electron microscope image of an L-IP steel showing austenite, ferrite
and κ-carbide [392].

In austenite-based L-IP steels, austenite serves the majority of phase, while ferrite and κ-
carbide are minority phases. Austenite-based L-IP steels possess excellent combination of
strength and ductility due to their high strain-hardening capacity. They are highly promising
candidates for automotive applications with the purpose of reducing energy consumption and
greenhouse gas emission [393-395]. An increase in aluminium content increases the volume
fraction of ferrite and decreases the mass density due to the smaller atomic weight of aluminium
compared to iron, as well as the difference in atomic density between the FCC and BCC
structures in steels [396]. It has been reported that the addition of 6.8 wt.% aluminium can lead
to a 7.5% density reduction in compared to conventional interstitial-free steel [397]. Ferrite-
based L-IP steels possess high work-hardening capacity due to the austenite-to-martensite
transformation during deformation, and thus to generate high strength and high ductility
lightweight steels through TRIP mechanism [398,399]. However, a high aluminium content
could result in the formation of DO3-type Fe3Al and B2-type FeAl intermetallic compounds,
which deteriorate the plasticity [400]. In order to achieve a better combination of strength and

69
ductility, TWIP effect can also be used through properly controlling the phase stability and
stacking fault energy of austenite. This is because deformation of austenite is dominated by
TRIP, TWIP and dislocation gliding, respectively, in an order with increasing the stacking fault
energy [401]. Formation of κ-carbide has a beneficial effect on the YS through precipitation
hardening. However, unfavourable morphology of κ-carbide may also affect the mechanical
properties [402]. To further improve the mechanical properties of ferrite-based L-IP steels, the
formation of κ-carbide can be suppressed through appropriate annealing treatment to make them
“austenite + ferrite” duplex steels [398]. Use of austenite as the second phase in the ferrite-
based L-IP steels can lead to an optimisation of strength and ductility through the combined
effects of TRIP and TWIP [403]. A work published in the journal Nature by Kim et al. [404]
shows that the hard intermetallic compound (B2) can be effectively used as a strengthening
second phase in high aluminium L-IP steels, while alleviating its harmful effect on ductility by
controlling its morphology and dispersion. It has been found that an increase in the aluminium
content to 6 wt.% can result in about a 10% weight reduction, compared to TRIP and TWIP
steels, and offers excellent properties, such as strength above 780 MPa and elongation above
30% [396].
However, aluminium reduces the work-hardening rate, resulting in relatively low fracture
strength [405]. It is also worth pointing out that L-IP steels offer processing challenges relative
to the other AHSSs. The complexity of the melting process of L-IP steels increases with the
addition of alloying element aluminium to a high content due to the large difference in the
specific gravity among iron (7.87 g/cm3), manganese (7.43 g/cm3), aluminium (2.70 g/cm3) and
carbon (2.25 g/cm3) [406]. Furthermore, aluminium increases the corrosion rate of lining
refractories (magnesia) of the furnace used in the melting process [407], which has the potential
to cause production accidents. Therefore, a careful consideration of both the required
mechanical behaviour and the production technology should be made in the design of the
chemical compositions of L-IP steels, especially the addition of aluminium element.
Recently, Chen et al. [408] reviewed the current state of Fe-Mn-Al-C low-density steels in
the aspects of physical metallurgy, processing strategies, strengthening mechanisms and
mechanical properties. Kim et al. [409] reviewed the microstructure and mechanical properties
of Fe-Al-Mn-C lightweight steels, focusing on the influence of alloying elements on phase
constituents and mechanical behaviour. Zuazo et al. [410] published a review article on the
complex metallurgy of low-density steels for automotive applications, in which the effects of
aluminium steel density, CALPHAD-type modelling, microstructure-property relationships and
automotive applications have been systematically described. However, there is still a lack of
review work on the TMP, especially processing-microstructure-property relationship, of L-IP

70
steels so far.
Similarly to other kinds of AHSSs, the TMP of L-IP steels also improves the performance of
the processed products through optimisation of processing parameters. However, TMP
schedules of L-IP steels are somewhat different due to their relatively high alloying contents,
particularly aluminium, which cause difficulties in processing. At present, it is still difficult to
do the TMP of L-IP steels in a single stage, and TMP is only a part of the processing schedule
of L-IP steels. Fig. 50 shows the typical flow diagram of the processing schedule of L-IP steels
on the basis of the work by Rana et al. [397,411]. The parameters, including reheating
temperature, rolling reduction, cooling method, coiling temperature, and annealing temperature,
applied in TMP are dependent on the steel’s chemical compositions, workability and the
targeted microstructure required for achieving improved properties.

Ingot casting Reheating Rough rolling Cooling

Coiling Cooling Hot rolling Reheating

Pickling Cold rolling Annealing Ageing

Warm rolling

Fig. 50. Typical flow diagram of L-IP steel processing.

The castability of L-IP steels is poor due to the loss of fluidity in the liquid steel and the
clogging problem of continuous casting [412]. Casting of L-IP steels is challenging and requires
technological innovation to improve the process. The purpose of rough rolling is to break down
the as-cast microstructure of L-IP steels. The reheating temperature for rough rolling generally
lies between 1200 and 1250 ℃, and the holding time may vary from one to several hours,
depending on the dimensions of steel ingots [399,413]. After rough rolling, the ingots are
generally cooled in air. Before hot rolling, steel slabs need to be reheated to between 1150 and
1250 ℃, then held at this temperature for some time to homogenise the microstructure,
depending on the steel’s chemical compositions and the thickness of the slabs [411,414,415].
Cracking is a serious problem that frequently occurs during hot rolling of L-IP steels, and is
greatly influenced by the microstructure of the steel plate and the presence of κ-carbides [398].
Shin et al. [416] suggested the following to prevent cracking in hot rolling of L-IP steels: (i)
reduce the volume fraction of κ-carbides formed between bands or along grain boundaries by
lowering the content of hardenability elements such as aluminium; (ii) stop the hot rolling

71
process before too many κ-carbides are formed; (iii) prevent the formation of κ-carbides that
continuously precipitate at band interfaces and decrease the volume fraction of the band
structure by increasing the rolling reduction ratio at high temperatures; and (iv) minimise the
central segregation during the slabmaking process and the material variation along the width
direction during hot rolling. In practice, strategies for preventing cracking of L-IP steels include:
decreased content of hardenability elements (e.g. aluminium); reduced central segregation
during the slabmaking process; decreased material variation during hot rolling; and relatively
high finish-rolling temperature. In order to prevent or minimise the occurrence of cracking in
hot rolling, the finish rolling temperature of L-IP steels is usually higher than 900 ℃, which is
above the κ-carbides dissolution starting and finishing temperatures (760 and 795 ℃) [398].
Also, flow instability should be considered in the design of hot rolling schedules, so that L-IP
steels are processed in the safe domain. The determination of flow instability, including
microband formation, crack initiation and flow localisation, of L-IP steels during hot
deformation can be realised with the aid of processing maps [417-419].
After hot rolling, the cooling rate from the finish rolling temperature to the coiling
temperature is important to determine the further processability and formability of L-IP steels.
The work of Sohn et al. [420] on ferrite-based L-IP steels showed that the steels should be
rapidly cooled from the finish rolling temperature to the coiling temperature to avoid the
lengthening or thickening of lamellar κ-carbides in order to prevent cracking during the cold
rolling process. Wang et al. [399] indicated that ferrite-based L-IP steels could be water-
quenched immediately after hot rolling to the coiling temperature to suppress κ-carbides
precipitation. The work of Rana et al. [421] showed that the application of a high cooling rate,
comparable to water quenching, after hot rolling to room temperature benefited the
improvement of formability of duplex L-IP steel in the final cold-rolled and annealed conditions
due to the suppression of κ-carbides precipitation.
Coiling temperature plays an important role in determining mechanical properties after
coiling. The selection of coiling temperature depends on the chemical composition of the steels
being processed. Generally, a high coiling temperature of 650 ℃ is used for ultralow-carbon
[397] or low-to-medium carbon [398,420,422-424] type L-IP steels. A low coiling temperature,
such as 400 ℃ [413], or even room temperature coiling [421], is applied to lean and rich alloy
L-IP steels. The lower coiling temperatures are used to facilitate further processing, such as cold
rolling, and improve formability in the final processed condition by minimising carbide
precipitation in the hot-rolled band [411]. Such a rule, however, may change for some specified
L-IP steels, e.g. a temperature of 550 ℃ was used to simulate the coiling process of a 0.95C-
27Mn-11.5Al-0.59Si duplex L-IP steel after hot rolling [425]. Wang et al. [415] conducted

72
research on the effects of coiling temperature ranging from 400 to 700 ℃ on the microstructure
and tensile properties of a hot-rolled 0.35C-1.1Mn-4.1Al-0.38Si L-IP steel, as schematically
shown in Fig. 51. They found that cracking often initiated at the interface between δ-ferrite and
original bainitic microstructure when coiling temperature was no higher than 450 ℃. For coiling
temperatures higher than 500 ℃, cracking could nucleate within carbide bands, at boundaries
between carbide bands and δ-ferrite bands, and/or at δ-ferrite grain boundaries within ferrite
bands. As a result, a coiling temperature of 450 ℃ was recommended to achieve a uniform
elongation of 25% and an elongation-to-failure of 32% through TRIP effect, due to the
formation of a higher volume fraction of lath-shaped retained austenite [399,415].

Fig. 51. Schematic illustration of hot rolling process of a 0.35C-1.1Mn-4.1Al-


0.38Si L-IP steel followed by coiling at different temperatures [415].

In order to remove the oxide scales from the surface, the coiled L-IP steel strip is usually
pickled in HCl solution at 85 ℃ [426]. Cold rolling after picking is needed to produce L-IP steel
at the final thickness. Through optimal control of the upstream processing variables, including
the cooling rate from the finish rolling temperature to the coiling temperature, a total reduction
of 67% could be applied to the cold rolling of L-IP steels containing aluminium less than 6
wt.% without cracking [398,420]. To achieve a much higher total reduction and prevent
cracking, cold rolling of L-IP steels can be combined with annealing prior to cold rolling
[427,428]. As aluminium raises the ductile-to-brittle transition temperature in steels, higher
aluminium-containing steels often crack during cold rolling, and therefore warm rolling or
“warm rolling + intermediate annealing + cold rolling” has been applied to reduce the
occurrence of cracking of L-IP steels with an aluminium content greater than 8 wt.%
[397,414,429].
Choosing the appropriate annealing temperature after cold or warm rolling is essential to
achieving the desired microstructure and targeted mechanical properties. The phase fraction in

73
L-IP steels is insensitive to annealing temperature. There is a broad range of temperatures that
can be used for annealing L-IP steels, depending on the final mechanical properties required
[430-436]. As tensile strength of L-IP steels generally increases at the expense of elongation, a
balance between strength and ductility should be taken into account before annealing treatment.
The work of Hwang et al. [427] on a duplex L-IP steel showed that ferrite grain size was nearly
unaltered but austenite grain size slightly increased with an increase in the annealing
temperature, revealing that there is little window to control the microstructure of the steel by
annealing. Sohn et al. [398] conducted research on the effect of annealing temperatures on the
microstructure and mechanical properties of a cold-rolled ferrite-based L-IP steel, as
schematically shown in Fig. 52. Their results indicated that the tensile strength increased, while
YS and elongation decreased when the annealing temperature was increased. Annealed at
830 ℃ or 880 ℃, a volume fraction of 22-24% fine austenite with appropriate mechanical
stability was formed and homogeneously distributed in the ferrite matrix, which led to the
obtaining of the best combination of strength and ductility. Park et al. [437] annealed cold-
rolled Fe-(8,12)Mn-5Al-0.2C L-IP steels at both 800 and 900 ℃, and found that YS annealed at
800 ℃ was higher than that at 900 ℃ due to insufficient annealing, and elongation showed an
increasing trend with the increase of temperature. For the Fe-8Mn-5Al-0.2C steel, a high tensile
strength of >900 MPa and a high total elongation of >50% could be obtained when the steel was
annealed at 900 ℃ due to the occurrence of deformation-induced martensitic transformation.
Haase et al. [438] suggested that the annealing temperature of L-IP steel should be high enough
to suppress κ-carbides precipitation, but also low enough to prevent extensive grain growth, so
that a balance between the strength and ductility can be achieved. Cold rolling and subsequent
annealing are essential processes of L-IP steels, and should be carefully considered in order to
achieve the required microstructure and targeted mechanical properties.

Fig. 52. Schematic illustration of the rolling and annealing conditions for a Fe-0.35C-3.5Mn-5.8Al
ferrite-base L-IP steel [398].
74
Ageing treatment after annealing is sometimes applied to L-IP steels to alter the state of
precipitates, and thus bring about improvements in mechanical properties. Therefore, in some
cases ageing is also an important part of processing. Song et al. [424] showed that the austenite
volume fraction in the ferrite-austenite duplex L-IP steel decreased as the ageing temperature
increased because some austenite grains decomposed to pearlites, and such austenite
decomposition to pearlite was favourable for the improvement of YS. Choi et al. [431]
performed ageing treatment on a Fe-28Mn-9Al-0.8C L-IP steel at 625 ℃ for different durations.
They found that extension of ageing time induced κ-carbides precipitation, and an increase in
the size and volume fraction of κ-carbides decreased the formation activity of slip bands,
resulting in a decrease in the work-hardening rate. The work of Lee et al. [439] indicated that
ageing could promote the precipitation of β-Mn in a Fe-31.4Mn-11.4Al-0.89C austenite-based
L-IP steel, and thus lead to increased hardness. The ageing temperature may change from 400 to
650 ℃, with the holding time varying from several to hundreds of hours, depending on the
chemical compositions of the steels being processed and the targeted mechanical properties
[413,424,431,439].

4. Thermomechanical processing of the third generation AHSSs


The first generation AHSSs are developed in fairly lean compositions, and provide higher
strength and better crash-worthiness than conventional high-strength steels. The high strength
and high elastic stiffness allow for a reduction in gauge thickness of materials and keep these
steels as the premier materials for automotive applications. However, the total elongation of this
generation AHSSs drops significantly with the increase of tensile strength, which deteriorates
the formability of the steels, and thus limits their applications in the automotive industry. The
second generation AHSSs clearly exhibit superior mechanical properties compared to both
conventional high-strength steels and the first generation AHSSs, due to the elevated alloying
additions, such as manganese, required to stabilise austenite at room temperature. However, the
second generation AHSSs are relatively expensive due to the high-cost alloying elements, which
is the main obstacle to their broad application. Also, industrial processing of these steels,
especially the TWIP steels with high manganese content, has shown to be extremely
challenging and the TWIP grades are also prone to delayed cracking [440]. Against this
background, the demand has grown for new third generation AHSSs that have high strength and
high formability, and are available at a reasonable cost.
The third generation AHSSs aim to produce steels with a better strength/ductility
combination than the first generation, without significantly enriching the alloy compositions, or
reducing the alloying levels of the second generation, with the aim of reducing production costs.

75
The targeted mechanical properties of the third generation AHSSs are intended to fall within the
gap between the first and the second generations, as indicated in Fig. 1. As the mechanical
properties of steels are primarily determined by the microstructural constituents, considerations
about microstructure are essential in the development of the third generation AHSSs. It has been
found that the mechanical properties of the third generation AHSSs can be achieved using steels
with austenite and martensite microstructures [441]. Steel with a microstructure that contains
austenite and martensite has much better strength/ductility combination than microstructures
that contain ferrite and martensite, and can be used to realise the mechanical properties required
for the third generation AHSSs. Fig. 53 shows the contributions of “austenite + martensite” and
“ferrite + martensite” microstructural combinations to the mechanical properties, in which the
dots on each curve indicate the drop in martensite volume fraction from the upper left to the
lower right. It is clear that the tensile strength increases with an increase in the martensite
volume fraction. It also shows that the volume fraction of austenite has a stronger influence on
ductility of the steel than that of ferrite. The mechanical properties of steels with a “ferrite +
martensite” microstructure overlap with the mechanical properties exhibited by the conventional
high-strength steels and the first generation AHSSs. The mechanical properties corresponding to
the “austenite + martensite” microstructures are situated between the mechanical properties of
the first and the second generations AHSSs, i.e. within the desired “third generation AHSS”
regime [442]. Based on Fig. 53, it is anticipated that microstructures of interest for the third
generation AHSSs will consist of high-volume fractions of metastable austenite with
capabilities to induce austenite-to-martensite transformation.

Fig. 53. Contributions of “austenite + martensite” and “ferrite + martensite”


microstructural combinations to the mechanical properties [442].

The stability of austenite has a significant effect on the mechanical properties. Fig. 54 shows
four hypothetical austenite stability curves based on Olson’s austenite stability function [443],

76
in which the cases A and D indicate the highest and lowest stabilities of austenite, respectively.
Table 2 lists the implications of the four cases exhibited in Fig. 54 [444]. It can be seen that the
lower the stability of austenite, the lower the strain that will lead to martensite formation. Fig.
55 shows the mechanical property combinations corresponding to ferrite plus austenite, with the
different stability cases of A, B, C and D [441]. In Fig. 55, the individual data points on each
curve correspond to different initial volume fractions of austenite, ranging from 0 to 85%, with
the remainder of the microstructure being ferrite. The lowest stability of austenite (case D) is
expected to present mechanical properties similar to the first generation AHSSs, indicating that
austenite does not significantly contribute to improved mechanical properties. As the stability of
austenite increases (cases C to A), new combinations of tensile strength and elongation can be
achieved. The best combination of tensile strength and elongation for high-volume fractions of
relatively stable austenite, in the case of B, can fulfil the requirement band of the mechanical
properties of the third generation AHSSs.

Fig. 54. Four different austenite stability cases identified as A, B, C and D [441].

Table 2 Implications of the four cases of A, B, C and D exhibited in Fig. 54 [444].


Olson and Cohen’s austenite stability function fα' = 1-exp{- β[1-exp(-αε)]n}
fα' = volume fraction of martensite
ε = plastic strain
Case α β N

A 2.64 1.55 2
B 5.5 5 2
C 12 5 2
D 50 5 2
Notes: α is a strain-independent constant, and represents the rate of shear-band formation; β is a temperature-
dependent parameter, and is proportional to the probability that an intersection will form an embryo, and this
probability is temperature dependent through its relation to the chemical driving force; and n is a fixed exponent.

77
Fig. 55. Mechanical property combinations corresponding to ferrite plus austenite with
different stability cases of A, B, C and D [441].

The stability of austenite that governs the austenite-to-martensite transformation is affected


by numerous factors, such as carbon content, austenite grain size, austenite morphology, and
steel’s chemical compostions. Considerations about these factors should be made when
processing the third generation AHSSs to optimise the microstructure and achieve the required
mechanical properties. A number of efforts have been made to obtain the third generation
AHSSs through methodological and technical developments [441,444-446]. It should be noted
that the information provided in Figs. 53 and 55 shows only an understanding of the
contributions of martensite, austenite and ferrite to the mechanical properties. In processing the
third generation AHSSs, complex microstructural constituents may be needed to obtain the best
combination of strength and ductility. The microstructure may include significant fractions of
high-strength phases such as martensite, bainite or ultrafine-grained ferrite, in combination with
highly-ductile austenite with controlled stability against austenite-to-martensite transformation
with strain [442,446].
In the last few years, a number of TMP technologies have been proposed to achieve the
mechanical properties required for the third generation AHSSs. Different to the first and second
generations, the TMP of the third generation is mainly focused on heat treatment because the
microstructural characteristics needed for the steels are primarily achieved by the control of
post-deformation heat treatment conditions, including annealing, cooling, partitioning and
tempering [447]. So far, only limited articles have addressed the rolling/deformation process in
the TMP of the third generation AHSSs. Askari-Paykani et al. [448] applied different
solidification routes and a two-step heat treatment process after hot rolling to a FeCrNiBSi alloy
system, and found that the mechanical properties of the alloy system could be adjusted within
either the first generation or the third generation through control of the cooling rate during

78
casting and heat treatment routes after hot rolling. De Moor et al. [449] summarised the results
of some processing technologies to generate high-strength austenite-containing microstructures
after adjustment to account for differences in specimen geometry, so that comparisons can be
made between the different investigations, as shown in Fig. 56. The solid and dashed lines are
the “ferrite + martensite” and “austenite + martensite” property predictions from Fig. 53
respectively. It is clear from Fig. 56 that several studies have yielded results plotting near or
above the dashed curve that is within the third generation AHSSs properties band. In this
section, five typical processing technologies, including Q&P, martensite-to-austenite reversion
treatment, quenching-partitioning-tempering (Q-P-T) and dual stabilisation heat treatment
(DSHT) and, that are attracting increasing interest for producing the third generation AHSSs,
will be reviewed.

Enhanced DP Quenching and partitioning


Ultrafine DP Jun-Fonstein
Modified TRIP Streicher et al.
Matsumura De Moor et al.
Micro-alloyed TRIP De Moor, Kwak, Lee et al.
TRIP Type Bainite Li et al.
TRIP-DUAL Wang et al. - ind. trial
Interrupted Quench Rapid Heating and Cooling
Bainite Flash process
Bhadeshia-Edmonds Lower Mn TWIP/TRIP
Miihkinen-Edmonds Frommeyer et al.
Caballero et al. Dastur-Leslie
Garcia-Mateo et al. Merwin

Fig. 56 Summary of combinations of elongation and tensile strength obtained by different


processing technologies [449].

4.1 Quenching and partitioning


The idea of Q&P was first proposed by Speer et al. [450] in 2003 for the development of the
third generation AHSSs, and it has been applied as a new way of producing martensitic steels
containing enhanced levels of austenite [451-453]. Q&P aims to create mixtures of carbon-
depleted martensite and carbon-enriched austenite through a controlled thermal treatment
process. A schematic of the Q&P process that enables carbon partitioning from martensite into
austenite is illustrated in Fig. 57 [442]. Q&P consists of a three-step thermal treatment. Steel is
firstly austenitised at the temperature higher than Ac3, followed by rapid cooling to a specific
quenching temperature to create a controlled volume fraction of martensite, followed by a
partitioning treatment at the partitioning temperature for a certain period to allow carbon
depletion of the martensite and carbon transport to the austenite, so as to obtain carbon-enriched

79
austenite. The partitioning step increases the stability of austenite and enables carbon-stabilised
austenite to be retained in the microstructure when the steel is finally quenched from
partitioning temperature to room temperature. Partitioning can be done at quenching
temperature (one-step Q&P), or by holding at a temperature higher than quenching temperature
(two-step partitioning) [454]. The Q&P process is applicable to the annealing of cold-rolled and
coated sheet steel products, whereas non-isothermal partitioning has been explored as a process
route for rolled products that typically encounter continuous cooling conditions after rolling
[454,455]. Q&P can also be applied to the hot stamping process of press-hardened steels, and
encouraging results have been obtained [456].

Fig. 57. Schematic illustration of the Q&P process for producing austenite-containing
microstructures stating from 100% austenite: Ci, Cγ and Cm represent the carbon contents
for the initial alloy, austenite and martensite, respectively [442]. QT: quenching
temperature, and PT: partitioning temperature.

In the Q&P process, the carbon partitioning from martensite into austenite is controlled by
the constrained carbon equilibrium criterion [457]. This criterion is to predict the carbon
concentration in austenite under three conditions, i.e. (i) an identical carbon chemical potential
exists in both ferrite (or martensite) and austenite; (ii) competing reactions, such as cementite or
transition carbide formation or bainite transformation, are suppressed; and (iii) the carbon
partitioning proceeds under the assumption that the interface between ferrite and austenite does
not migrate. However, this model does not account for the volume expansion that usually occurs
during the partitioning step. Volume expansion in Q&P is caused by the bainitic transformation
or the migration of martensite/austenite interface when the partitioning temperature is above the
Ms temperature, and may contribute to the enrichment of carbon into austenite [458]. In early
Q&P steels, silicon and/or aluminium are usually added to suppress carbide formation to ensure
the best effect of austenite stabilisation, and consequently, the carbide formation elements such
as niobium, molybdenum and vanadium, are eliminated from Q&P steels [459]. As a result,

80
carbon is enriched in the remaining austenite after quenching during the partitioning process,
and a certain amount of austenite is retained after the final cooling to room temperature [460].
However, carbide precipitation in martensite takes place frequently in low-carbon and high-
carbon steels, even though they contain a high amount of silicon [461]. Some of the carbon is
consumed to form carbide due to carbide precipitates, which reduce the remaining amount of
carbon in martensite that can be enriched in austenite during partitioning. In this case, the
carbon concentration in austenite after partitioning will be lower than predicted under the
constrained carbon equilibrium conditions, excluding carbide precipitation.
Carbide precipitation in the Q&P process may also have positive effect on mechanical
properties if the size and distribution of precipitates can be properly controlled. This control can
be realised by a modified Q&P process, i.e. Q-P-T treatment, which is discussed in Section 4.3.
Q&P is mainly applied to the steels with chemical compositions similar to those of TRIP steels.
In these steels, bainite is generally unavoidably formed because their chemical compositions are
designed to promote bainite formation during tempering in the same temperature range as the
partitioning step [458]. During tempering, the carbide precipitation and bainitic transformation
greatly affect carbon enrichment in untransformed austenite because both microstructural
changes consume carbon and further affect the amount of stability of austenite. Therefore, a
proper quenching temperature should be applied in the Q-P-T process to achieve the maximum
fraction of austenite. The value of quenching temperature can be determined according to the
constrained carbon equilibrium theory proposed by Speer et al. [457] combined with the
Koistinen-Marburger kinetics [462] of martensitic transformation. The determined tempering
temperature should benefit the formation of fine carbides that are dispersed in the matrix and
satisfy the requirement of carbon partitioning from martensite to austenite [463,464].
In the Q&P process, carbon partitioning and microstructural evolution are decoupled. The
microstructures are controlled by an athermal martensitic transformation, while carbon
partitioning is controlled in an isothermal stage [465]. The microstructural evolution in the Q&P
process is more complex than that with only martensite formation followed by carbon
partitioning from martensite to austenite. The work of Santofimia et al. [466] indicated that the
morphology of the resulting microstructure after Q&P treatment not only depends on the
parameters of the heat treatment, but also on the initial microstructure prior to this process. It is
therefore possible to create different final microstructures by using different initial
microstructures prior to the Q&P process that may generate new mechanical property
combinations. The traditional Q&P process starts with reheating the steels to a temperature
higher than Ac3 to generate fully austenitised microstructure. To create variable initial
microstructures before Q&P, the steels can be treated by a partial austenitisation process to

81
produce steels with a microstructure composed of ferrite, carbon-depleted martensite, and
carbon-enriched retained austenite [467]. The selection of either full austenitisation or partial
austenitisation prior to Q&P depends on the targeted mechanical properties [468].
A Q&P process starting with full austenitisation is generally used on steels for both bar and
sheet applications. Whereas steels that are processed starting with partial austenitisation, and
which contain a certain amount of ferrite in the microstructure, are mainly cold rolled for sheet
applications [468]. The microstructure generated after partial austenitisation can lead to an
excellent combination of mechanical properties, in which the good formability is attributed to
the TRIP effect from the retained austenite and ferrite, and the enhanced strength is due to the
presence of martensite instead of bainite. The work of Santofimia [469] showed that the
presence of a high fraction of initial martensite promoted the formation of film-like Q&P
microstructures, whereas a predominant presence of allotriomorphic ferrite in the initial
microstructure led to polygonal Q&P morphologies. Zhang et al. [467] found that a partially
austenitised Q&P treatment could significantly enhance the product’s strength, with elongation
of a Nb-microallyed low-carbon steel from 16,000 MPa% to more than 24,000 MPa%,
compared to a fully austenitised Q&P process.
Fig. 58 compares the combinations of UTS and total elongation of the third generation
AHSSs produced by Q&P processing with that of TRIP, DP and MART steels [441]. The data
in Fig. 58 indicate that Q&P is a promising method to produce the third generation AHSSs with
improved strength and ductility. However, the preferred times and temperatures for partitioning
treatment, and the optimal final quenching temperature to yield the maximum amount of
austenite have not been clearly understood [470]. At present, most work on Q&P focuses on the
hot-rolling off-line heat treatment. There is still no report available on Q&P at a practical hot-
rolling on-line industrial scale, even though direct Q&P has been successfully applied after hot
rolling in a laboratory environment [471]. Significantly more work is required to optimise the
steel compositions and Q&P processes with the purpose of producing new steel grades that can
fulfil the third generation AHSSs properties band, and to realise transition from scientific
research to industrial production.

82
Fig. 58. Total elongation vs. UTS for sheet steels processed with microstructural
characteristics of TRIP, DP, MART and Q&P steels [441].

4.2 Martensite-to-austenite reversion treatment


Q&P steels are stronger but less ductile when compared to TRIP steels. The lower ductility
is a result of the lower volume fraction of austenite, and this is associated with the TRIP effect
[472,473]. If the stability and volume fraction of austenite can be well controlled without a
significant decrease in strength, the mechanical properties of Q&P steels can be further
improved. It has been well recognised that a high amount of stable austenite in TWIP steels can
be achieved by alloying with substitutional austenite stabilisers [444]. Due to the high cost of
TWIP steels, researchers have been working on new processing methods of low-alloying steels
with a purpose of retaining a high amount of stable austenite at an acceptable cost level. Lee et
al. [474] summarised that three main factors, including chemical compositions [475], grain size
[476] and mechanical stabilities [477], dominated the stability of metastable austenite at room
temperature. These factors should be taken into account in an effective processing design.
Martensite-to-austenite reversion treatment, which involves reversion of austenite from
martensitic microstructure during the annealing process, is beneficial in increasing the volume
fraction of austenite due to the consequence of redistribution of alloying elements, and
enhancing the stability of the treated austenite. Several studies have indicated that reversion-
treated steel can achieve an outstanding tensile strength of about 1-1.5 GPa and a total
elongation of about 31%-44% after obtaining a high volume fraction of about 30% austenite by
enrichment of carbon and manganese [478,479].
Martensite-to-austenite reversion treatment is commonly performed on medium manganese

83
steels. However, it can also be performed on steels with low manganese additions, depending on
the targeted mechanical properties [480,481]. Medium manganese steels are regarded as the
third generation AHSSs due to their excellent mechanical properties and relatively low cost
(compared to high manganese steels). Hu et al. [482] recently reviewed the design strategies of
medium manganese steels with the aim of achieving similar tensile properties to the classical
TWIP steels that contain higher manganese content. As the mechanical properties of medium
manganese steels depend strongly on the characteristics of retained austenite, including its
amount, grain size, morphology and stability, the control of retained austenite becomes crucial.
During the martensite-to-austenite reversion process, the reverted austenite is primarily
stabilised by manganese and nickel [483]. However, due to the high cost of nickel, much
attention has been paid on the manganese-containing steels with reduced or no nickel additions
for the martensite-to-austenite reversion treatment [444,474,478,483]. In the martensite-to-
austenite reversion treatment, the reversion process leads to segregation of manganese and
carbon to the grain boundaries upon annealing, which promotes the nucleation of nanolaminate
austenite at martensite grain boundaries [484,485]. A significant degree of manganese and
carbon partition into the reverted austenite grains occurs in the martensite-to-austenite reversion
process, which helps stabilise the reverted austenite at room temperature because both
manganese and carbon are austenite stabilisers [486-488].
The amount, grain size and morphology of reverted austenite in medium manganese steels
are dependent on the annealing conditions [489]. At relatively low annealing temperatures, the
reverted austenite possesses good thermal stability, due to the significant amount of manganese
partitioning, and can be retained at room temperature even upon quenching. However, the
kinetics of austenite reversion at the low annealing temperature is very slow, and it may take
hours or even tens of hours to obtain a certain amount of reverted austenite, which limits
practical applications in industry [490, 491]. By increasing the annealing temperature, the
kinetics of austenite reversion and the amount of reverted austenite under the same holding time
would be increased due to the increased atomic diffusion coefficients of both manganese and
carbon. The optimal annealing temperature needed to achieve the maximum amount of austenite
is dependent on the steel’s chemical compositions. For cold-rolled TRIP-assisted steels, the
optimum annealing temperature for increasing the volume fraction of reverted austenite can be
estimated by “(Ac1 + Ac3)/2 + 20 ℃” [481]. An increased amount of reverted austenite is
essential for obtaining excellent combination of high YS, good elongation and excellent
toughness [478]. However, as the stability of the reverted austenite decreases with the increased
annealing temperature, some of the reverted austenite with insufficient stability is transformed
back to martensite during final cooling to room temperature [492]. A higher annealing

84
temperature induces a greater partition of manganese and carbon to austenite [493]. The work of
Lee et al. [474] on a Fe-0.05C-6.15Mn-1.4Si-0.04Al medium manganese steel indicated that
very efficient manganese partitioning into austenite occurred during a 180 s annealing in the
temperature range of 640-680 ℃. Luo et al. [478] found that a significant enrichment of
manganese in the austenite lath could be obtained after a 5 min annealing. An enrichment of
manganese in austenite significantly improves the elongation and lowers the YS, but has less
effect on the UTS of the intercritically annealed steels [478].
The work of Zeytin et al. [494] showed that the holding time for annealing had different
effects on the formation of retained austenite in different steels after annealing treatment. For a
specified steel, the volume fraction of austenite increases first, then it decreases after a peak
value with the increase of annealing holding time [495]. Arlazarov et al. [496] conducted
annealing on a medium manganese steel and found that the amount of reverted austenite could
achieve ~15% after 1 h of holding, and it reached the maximum value of ~22% when the
holding time was increased to 7 h. A further increase in the holding time induced a fall of the
volume fraction of austenite, and only a few percent (~6%) of austenite could be obtained when
the holding time was increased as long as 30 h. Huang et al. [497] also found that the amount of
reverted austenite was virtually insensitive to the cooling rate after annealing. Also, the amount
of austenite passed through a peak value, after which the amount of austenite dropped with an
increase in the holding time.
The work of Wang et al. [486] on a 9Mn medium manganese steel showed that a thin layer
of reverted nanoscale austenite grows along the prior austenite grain boundaries and martensite
packet/block boundaries after annealing at 600 ℃ for 1 h. After 8 h further annealing, all
martensite lath boundaries are decorated by reverted nanoscale austenite grains. The extended
annealing treatment leads to a quasi-continuous reverted nanoscale austenite network that
decorates all martensitic boundaries with only limited gaps in between. As a result, a consistent
improvement in both ductility and toughness was achieved, with only limited sacrifice of YS or
UTS, when the annealing time was increased from 1 to 8 h. In the martensite-to-austenite
reversion process, long time annealing will lead to excessive manganese and carbon enrichment
in austenite, which renders austenite too stable and therefore weakens the TRIP effect.
Furthermore, long time annealing reduces dislocation density, induces coarsening of nano-sized
precipitates, and causes a low work-hardening rate of the treated steel [498-501]. As a result, the
YS and UTS of medium manganese steels would deteriorate if the annealing time is too long
[502,503]. However, the annealing time should also not be too short because the microstructure
required for the improvement of mechanical properties may not be properly formed during a
very short annealing time. Hu et al. [504] found that the elongation of a 5Mn steel was

85
significantly increased from 8.9% to 25.0% when the annealing time was increased from 10 to
30 min at the same annealing temperature of 600 ℃, and also, the tensile strength was increased
from 829 to 875 MPa. The impact energy at -60 ℃ was significantly increased from 84.2 to
165.3 J. This is due to the degree of cementite dissolution, the recovery of martensitic structure,
and the fraction of reverted austenite with appropriate mechanical stability. These key factors
contribute to the increased strength, ductility and toughness of the steel, and are greatly
improved when the annealing time is increased from 10 to 30 min.
Table 3 presents some examples of mechanical properties of medium manganese steels
which suffered from martensite-to-austenite reversion treatment under different annealing
conditions. It can be seen that the mechanical properties of medium manganese steels are
greatly affected by the annealing temperature and the holding time. The mechanical properties
are different for different steel grades, even though the annealing condition is the same. For
example, when annealing at 650 ℃ for 10 min, the YS, UTS and elongation are respectively 645
MPa, 751 MPa and 31.6% for the steel 0.04C-5Mn-0.2Si-0.2Mo, but the values become 769
MPa, 829 MPa and 8.9%, respectively, for the steel 0.1C-5Mn-0.2Si-0.4Mo. Therefore,
different strategies of martensite-to-austenite reversion treatment should be made for different
grades of medium manganese steels with the purposes of obtaining optimal microstructure and
achieving improved mechanical properties.

Table 3 Mechanical properties of medium manganese steels which suffered from


martensite-to-austenite reversion treatment under different annealing conditions.
Steels Temperature Holding time YS UTS Elongation Reference
0.04C-5Mn-0.2Si-0.2Mo 600 ℃ 10 min 725 MPa 776 MPa 27.8% [491]
650 ℃ 10 min 645 MPa 751 MPa 31.6%
0.035C-5.1Mn-0.2Si-1.4Ni 600 ℃ 2h 735 MPa 787 MPa 20% [492]
650 ℃ 2h 588 MPa 915 MPa 14.5%
0.1C-5Mn-0.2Si-0.4Mo 650 ℃ 10 min 769 MPa 829 MPa 8.9% [504]
650 ℃ 30 min 770 MPa 875 MPa 25%
0.01C-9Mn-3Ni-1.4Al 600 ℃ 1h 760 MPa 920 MPa 9.9% [486]
600 ℃ 8h 665 MPa 900 MPa 17.1%

In order to improve the ductile without sacrificing the strength especially UTS, an effective
approach is to further refine the microstructure and stabilise the reverted austenite through
optimal martensite-to-austenite treatment on the basis of the annealing process. Pre-quenching
prior to annealing can be used to refine the microstructure and increase the volume fraction of
stable reverted austenite, which is beneficial for the improvement of both strength and
elongation [505]. The phrase “pre-quenching” means steel is first quenched from the
austenitising temperature to the room temperature prior to the annealing process. The purpose of
pre-quenching is to generate a fine lath martensite initial microstructure, which promotes the
formation of refined and stable reverted austenite after the following annealing treatment. The

86
lath martensite offers a large amount of nucleation sites that form ultrafine austenite laths after
annealing, and remove the previously formed carbides that may deteriorate the stability of
reverted austenite. As only a short annealing time is required, the “pre-quenching + annealing”
process has potential for the massive production of medium manganese steels in industrial
production lines [493,506].
Zhu et al. [507] reported a cyclic martensite-to-austenite reversion treatment with the aim of
stabilising the reverted austenite in a 0.21C-4.53Mn medium manganese steel. In this approach,
steel was first annealed at 700 ℃ for 10 min before quenching to room temperature. Then, the
thermal cycle was repeated twice, so that the total holding time at 700 ℃ was 30 min. In the first
cycle of treatment, some of the reverted austenite transforms back to martensite after quenching
to room temperature. Then in the second cycle, carbon in supersaturated martensite partitions to
austenite nearby, thereby increasing the stability of reverted austenite. Also, the fresh martensite
formed during the first cycle possesses a large amount of grain boundaries and defects, which
act as preferential nucleation sites for the newly formed austenite. As a result, the stability of
reverted austenite after the second cycle treatment is increased. Such mechanism works
continuously in the third cycle. Finally, the reverted austenite is effectively stabilised and the
amount of reverted austenite is greatly increased after cyclic reversion treatment.
Medium manganese steels are commonly quenched to room temperature before reversion
treatment. For improving the mechanical properties of medium manganese steel, Tsuchiyama et
al. [487] modified the reversion treatment process by introducing interrupted quenching. In this
method, the steel was first quenched to a temperature between Ms and Mf for a period of time to
retain a certain amount of austenite, then heated up to the annealing temperature to allow
martensite-to-austenite reversion. This process can generate a unique multi-phase structure
consisting of tempered martensite, fresh martensite, and two types of granular and film-like
retained austenite. Compared to the quenched and annealed specimen, the specimen treated by
interrupted quenching and annealing has much higher tensile strength (1070 MPa), while still
possessing adequate ductility (elongation: 33%) due to the combined effects of fresh martensite
and TRIP during tensile deformation. Wang et al. [508] found that the cold-rolled martensite
can provide a higher density of martensite grain boundaries compared to as-quenched martensite,
which benefits the nucleation and growth of reverted austenite grains in the subsequent
annealing process. The resulting microstructure consists of martensite and reverted austenite
grains with a well-dispersed size distribution, which promotes the TRIP effect over a wide
regime of strain, and contributes to the improvement of both strength and ductility.
The improvement in mechanical properties in the reversion-treated medium manganese
steels is associated with the high mechanical stability introduced by the TRIP/TWIP-induced

87
dynamic strain partitioning and nanolaminate morphology-enabled damage resistance [486]. In
the early uniform deformation regime, the reverted nanoscale austenite phase partitions the
majority of the strain via formation of twins, stacking faults and martensitic phase
transformation. At these small strains, only certain martensitic regions with preferred
orientations are plastically deformed. At late uniform deformation regime, more martensitic
regions become plastic and a co-strain partitioning process of both the reverted nanoscale
austenite and the martensite takes place. During further straining up to the late uniform
deformation regime, the martensitic phase transformation is completed and all martensitic
regions are rendered plastic. In the post-necking regime, the damage process takes place via
nucleation and coalescence of nanovoids, but most incidents stay confined by this inherited
nanolaminate microstructure morphology [509,510].
The size of reverted nanoscale austenite grains influences the deformation mechanism of
medium manganese steels. In coarser reverted nanoscale austenite grains, mechanical twinning
is favoured and leads to the build-up of complex and dense in-grain deformation structures. The
development of in-grain deformation substructures stabilises the grains and results in higher
phase stability of coarser reverted nanoscale austenite grains. For finer reverted nanoscale
austenite grains, mechanical twinning is less favoured. Less pronounced in-grain deformation
substructures form in these grains. Hence, they transform into martensite even at small
deformations. As a result, finer grains exhibit lower stability against transformation [511]. The
size effects of reverted nanoscale austenite grains can be used to modify the deformation
mechanisms of medium manganese steels at different stages of deformation in combination with
strain partitioning, thus improving the resistance to strain localisation and failure of the steels.
A number of research works have shown that martensite-to-austenite reversion treatment can
be used to simultaneously improve the strength, ductility and impact toughness of medium
manganese steels if optimal treatment schedule is applied [486,491,492,504,511]. Recently,
Koyama et al. [512] reported that the fatigue properties of medium manganese steel can also be
improved through controlling the reverted austenite from martensite. In their research, the
reversion-treated steels possess a key microstructural characteristic of multi-phase, metastable
and nanolaminate, which contribute to the significant improvement of fatigue performance of
medium manganese steels through simultaneous introduction of roughness-induced crack
termination and transformation-induced crack termination mechanisms, as compared to pearlitic,
TRIP and DP steels. As a TMP technology for medium manganese steels, martensite-to-
austenite reversion treatment provides an effective way to modify the microstructure and
improve the mechanical properties by a choice of appropriate TMP routes.

88
The key to martensite-to-austenite reversion treatment is to control the state of reverted
austenite, including its amount, grain size, morphology and stability, because these factors
significantly affect the mechanical properties of medium manganese steels, and thus determine
the effectiveness of reversion treatment schedules. In order to obtain the desired microstructural
constituents and achieve the improved mechanical properties, the state of martensite before
reversion treatment, as well as the annealing conditions, should be carefully considered in the
design of martensite-to-austenite reversion treatment strategies. As the promising materials for
automotive applications, medium manganese steels with improved strength, ductility, impact
toughness and fatigue properties are achievable through optimal design of TMP schedules and
appropriate control of martensite-to-austenite reversion process.

4.3 Quenching-partitioning-tempering
Q-P-T was developed as a modified Q&P process by adding some carbide-forming elements,
such as niobium and/or molybdenum, in the treated steels [513]. The Q-P-T treated steels
should contain less than 0.5 wt.% carbon to avoid the formation of cementite which may cause
quench and temper embrittlement [514]. In the Q-P-T process, steel is first quenched from a
proper austenitising temperature to a suitable quenching temperature, between Ms and Mf, to
obtain the maximum fraction of retained austenite. In the following partitioning and tempering,
carbon atoms partition from supersaturated martensite into nearby austenite and precipitation of
nano-sized carbides from martensite matrix, and the former leads to the stability of carbon-
enriched retained austenite at room temperature [459,515]. The Q-P-T process introduces
tempering so that fine carbides can be precipitated during tempering step to dig out the effect of
precipitation strengthening in the steels [516,517]. During tempering, carbon enrichment in
untransformed austenite is closely related to carbide precipitation and bainitic transformation,
since these microstructure changes also consume carbon and further influence the amount and
stability of retained austenite [516].
Zhong et al. [36] found that a good combination of tensile strength (1500 MPa) and
elongation (15%) could be obtained in a Fe–0.2C–1.5Mn–1.5Si–0.05Nb–0.13Mo steel
subjected to Q–P–T process. The work of Zhang et al. [440] showed that the product of strength
and elongation of a Fe–0.42C–1.46Mn–1.58Si–0.028Nb steel could be achieved as high as
31,627 MPa%, which is much higher than that processed by traditional quenching and
tempering. Extensive studies have indicated that Fe-Mn-Si-Nb alloyed steels subjected to Q-P-T
exhibit a better combination of strength and elongation compared to conventional Q&P steels
[36,459,463,518]. The mechanical properties of DP, TRIP, martensitic, Q&P and Q-P-T treated
steels are compared in Fig. 59 [514]. It can be seen that the Q-P-T steels exhibit a better

89
combination of strength and ductility than the Q&P steels and other grades of AHSSs. The Q-P-
T steels have become a new family of AHSSs with ultra-high strength associated with
considerable toughness and ductility [36,463,514,515,517,519].

Fig. 59. Total elongation versus UTS for the DP, TRIP, martensitic (M), Q&P and
Q-P-T steels [514].

4.4 Dual stabilisation heat treatment


Qu et al. [520] proposed the DSHT concept with the purpose of increasing the content and
stability of retained austenite, and therefore improving the combined mechanical properties of
the third generation AHSSs. DSHT is a variant of the Q&P concept that is compatible with
current methods of continuous production of galvanised steel. Five stages, including
austenitisation, initial quench, final quench, carbon partitioning and air cooling, are involved in
the DSHT process, as shown in Fig. 60 [444]. In the DSHT process, steel is first quenched after
austenisation to a temperature above the martensitic transformation starting temperature, Ms, to
mimic the thermal history used in galvanised steel production. The steel is further quenched to a
temperature below Ms to allow a controlled degree of martensitic transformation. This second
quench is followed by a carbon-partitioning step, with final air cooling to room temperature
[521]. DSHT can generate a three-phase microstructure, including ferrite, martensite and
retained austenite, with refined austenite grains. The resulting microstructure after DSHT
process provides a high tensile strength up to 1650 MPa combined with elongation of about
20% that satisfy the demands for the third generation AHSSs [520]. However, one problem with
this process is that the processing steps of the DSHT treatment are complex for applications in
industry [522].

90
Fig. 60. A schematic representation of the DSHT process [444].

5. Thermomechancal processing of Nano Hiten steels


The excellent strength-ductility balance of the three (first, second and third) generations of
AHSSs is achieved through introducing different combinations of phases, such as ferrite,
martensite, bainite and retained austenite, into the final microstructure. However, these
multiphase steels often possess poor hole expansionability, which has hindered their wide
application in the automotive industry. The poor hole expansionability of these AHSSs, as
represented by DP steels, comes from the large difference in deformation capabilities of the soft
and hard phases, and microvoids are commonly generated at the interphase between these
phases during punching. In order to overcome this problem, steels with single-phase structures,
such as bainite or bainitic ferrite, have been developed. However, the elongation of bainitic or
bainitic steels is generally low. With multiphase, such as “ferrite + martensite”, high elongation
can be obtained, but the hole expansionability is low. However, an improved hole expansion
ratio along with high elongation can be achieved by adopting a single ferrite phase
microstructure, as indicated in Fig. 61 [523].

Fig. 61. Effect of microstructure on elongation and hole expansion ratio [523].

Nano Hiten steels have a ferrite matrix strengthened by a large amount of nano-sized
precipitates. In the absence of peralite and coarese cementite, this group of ferritic steels can
91
achieve improved hole expansionability while maintaining high elongation, due to a lower level
of stress-concentrating regions compared to the steels with a multiphase microstructure, such as
DP steels. Nano Hiten steels have three key features [14,524]:
(1) single ferrite phase microstructure with excellent formability;
(2) strengthening by nano-sized precipitates; and
(3) extremly high thermal stability of precipitates.
Precipitates in steels can be classified into three categories based on their different nucleation
stages: (i) precipitates formed in austenite; (ii) precipitates formed at the austenite/ferrite
interface during the austenite-to-ferrite phase transformation; and (iii) precipitates formed in
supersaturated ferrite. Nano Hiten steels take advantage of interface precipitates. These
precipitates can be significantly refined to serve as highly effective strength enhancers through
dispersion strengthening, compared to the coarse precipitates formed in austenite or
supersaturated ferrite [525,526]. The degree of strengthening by interface precipitates is
dependent upon their type, size, number density and thermal stability. It is generally accepted
that the interface precipitate (Ti,Mo)C in Nano Hiten steels possesses a NaCl-type crystal
structure with a lattice parameter of 0.433 nm [525,527,528], even though a Ti0.98Mo0.02C0.6 type
precipitate with a hexagonal lattice and lattice parameters of a = 0.31 nm and c = 1.5 nm was
identified in a Ti-Mo bearing Nano Hiten steel [529].
(Ti,Mo)C can strongly maintain nano-scale size and makes the largest contribution to the
steel’s strengthening, compared to other kinds of precipitates such as TiC and (Ti,Nb)C [530].
The work of Shimizu et al. [14] indicates that the size of precipitates should be less than 10 nm
and the number density should be high enough that the role of precipitation strengthening in
Nano Hiten steels can be well played. Thermal stability of precipitates is also an important
factor to be considered because coarsening can easily take place if precipitates are fine but
thermally unstable, due to variations in processing conditions. Seto et al. [531] compared a
Nano Hiten steel strengthened by (Ti,Mo)C with a conventional HSLA steel strengthened by
TiC in term of tensile strength, as shown in Fig. 62. A large drop in tensile strength associated
with coarsening of TiC can be observed in the conventional HSLA steel when it was held for 15
× 103 s. For Nano Hiten steel, there is virtually no decrease in tensile strength even after holding
for 80 × 103 s if precipitation strengthening is achieved with (Ti,Mo)C. This result confirms that
the (Ti,Mo)C precipitates in Nano Hiten steel have extremely high thermal stability. The
advantages of (Ti,Mo)C precipitates allow them to be used as an excellent precipitation
strengthening provider, and this explains why most Nano Hiten steels are alloyed with titanium
and molybdenum to form refined (Ti,Mo)C in processing with the aim of achieving maximum
precipitation strengthening [14,527,531-534].

92
Fig. 62. Comparison of the thermal stability of tensile strength between Nano Hiten and
conventional HSLA steels [531]. Note: both steels were reheated to 923 K followed by
isothermal holding with different times.

The pearlitic transformation and coarsening of cementite easily occur when conventional
HSLA steels are coiled at high temperatures. As these constituents are detrimental to the
toughness of steels, their formation should be restricted. In Nano Hiten steels, a single-phase
ferrite structure can be obtained because the carbon content is low, and molybdenum is added to
retard the deposition of both pearlite and coarse cementite at grain boundaries, thus preventing
pearlitic transformation [523]. Manganese prevents fine precipitates from coarsening by
lowering the Ar3 temperature. As nitrogen may form TiN and thus exhaust titanium, the nitrogen
content in Nano Hiten steels should be controlled at a low level. A typical research study
conducted by Funakawa et al. [525] indicated that both the UTS and YS of Nano Hiten steels
exhibited the maximum values when the atomic concentration ratio of Ti/Mo was kept at 1.
Beyond the peaks, the tensile strength slightly decreased with the increase in the Ti/Mo atomic
concentration ratio. The elongation of Nano Hiten steels decreased unvaryingly with the
increase in the Ti/Mo atomic concentration ratio. Thus, the contents of titanium and
molybdenum should be carefully considered in the design of steel’s chemical compositions
based on the requirement of targeted mechanical properties.
During the autenite-to-ferrite transformation, precipitates nucleate preferably at the
austenite/ferrite interface because both strain and surface energy can be minimised with
precipitation. As the migration of the austenite/ferrite interphase boundary is driven by the
concentration gradient of carbon in the adjacent austenite, only carbon partitioning takes place.
There is also no movement of metallic solutes across the interface boundary when the
movement of the interphase boundary pauses for precipitation nucleation. The growth of
precipitates occurs rapidly because of the easy diffusion of metallic solute atoms along the
93
interphase boundary. Following the formation of interphase precipitates, the carbon
concentration ahead of the austenite/ferrite interphase becomes favourable and the interphase
boundary migrates to a new position where the nucleation cycle occurs again. This precipitation
process can be repeated many times during the austenite-to-ferrite phase transformation, thereby
leading to a very fine row-like dispersion of carbides in the ferrite matrix [525,528,532,535].
The smaller the spacing of rows and interparticles, the higher the precipitation strengthening due
to the increased number density of precipitates in the ferrite matrix.
The TMP of Nano Hiten steels is similar to that of ordinary steels, and a precipitation-
strengthened single ferrite phase can be easily obtained at a proper coiling temperature.
However, nano-sized precipitates in Nano Hiten steels do not always appear due to improper
TMP treatment. A TMP schedule with optimal processing parameters is therefore essential to
maximise precipitation strengthening by refining the interface preicipitates to a size of several
nanometres. In general, the TMP of a Nano Hiten steel includes the following five key stages: (i)
reheating the steel to a temperature above the dissolution temperature of carbides and
carbonitrides, and helding for some time; (ii) performing deformation operations above the Ar1
temperature; (iii) cooling the deformed steel to the coiling temperature under a high cooling rate;
(iv) holding the steel at the isothermal temperature for some time; and (v) cooling the processed
steel slowly to room temperature. In this section, the TMP of Nano Hiten steels is reviewed in
the sequence of the processing stages (i), (ii), (iii), (iv) and (v), as schematically shown in Fig.
63.

i
Temperature

ii

Ar1
iii
iv

Isothermal temperature
v

Time
Fig. 63. Schematic illustration of TMP schedule of Nano Hiten steels.

In the TMP of Nano Hiten steels, the purpose of reheating is to dissolve the carbides and
carbonitrides, and homogenise the austenite microstructure. Therefore, the reheating
94
temperature should be high enough to dissolve the carbides and carbonitrides. In practice, Nano
Hiten steels are commonly reheated to 1200 ℃ [527,528,532,534]. At this temperature, most
carbides, including VC, VN, NbC, NbN and TiC, will dissolve because their dissolution
temperatures are 777, 974, 1095, 1114 and 1068 ℃, respectively [44]. However, Nb(C,N) could
remain as the temperature for completely dissolving this carbide is higher than 1200 ℃ [536].
As Nb(C,N) can be completely dissolved at the reheating temperature of 1250 ℃ [44], a large
number of research studies have applied this temperature in the TMP of Nano Hiten steels
[525,533,537-542]. Other researchers have used temperatures that are higher than 1250 ℃, e.g.
1280 ℃ [543] and 1300 ℃ [22], to dissolve the precipitates more sufficiently, given the
possibility that niobium is not fully soluble at the reheating temperature of 1250 ℃, and could
remain in titanium-rich (Ti,Nb)(C,N) carbonitride [183].
Typically, there are three regions in the TMP of Nano Hiten steels in which deformation can
be processed. These regions are: austenite recrystallisation; austenite non-recrystallisation; and
“austenite + ferrite” dual-phase, with deformation temperatures varying from high to low. Kim
et al. [22] found that samples rolled in the austenite non-recrystallisation region are
characterised by finer ferrite grains but coarser precipitates, compared to those rolled in the
austenite recrystallisation region. One reason is that finer austenite grain size can be obtained
when steel is processed in the austenite non-recrystallisation region. Ferrite grain size is closely
related to the austenite grain size because austenite grain boundaries serve as the nucleation sites
of ferrite during the austenite-to-ferrite transformation process [64]. Also, a mass of dislocations
are accumulated in the samples rolled in the austenite non-recrystallisation region due to the
difficulties in accommodating deformation. As the dislocations can also serve as the nucleation
sites of ferrite, the ferrite grain size after austenite-to-ferrite transformation will be further
refined. However, rolling in the austenite non-recrystallisation region can bring about strain-
induced precipitation due to the presence of accumulated dislocations in austenite grains [544].
The precipitates grow during the subsequent processes. The particles precipitated in austenite
non-recrystallisation region and their subsequent growth consume the carbide formers of
titanium, molybdenum and carbon, and as a result, the number density of fine precipitates
formed during the following isothermal treatment process is decreased.
In order to balance the size of ferrite grains and precipitates, and therefore achieve an
optimised TMP schedule, Kim et al. [545] systematically studied five rolling conditions across
the regions of austenite recrystallisation, austenite non-recrystallisation and “austenite + ferrite”
dual-phase, as schematically shown in Fig. 64. Their results indicate that a much more effective
grain refinement can be achieved by conducting rolling in the regions through austenite
recrystallisation to austenite non-recrystallisation, i.e. the rolling conditions ② and ③. Under

95
rolling condition ② , because most of rolling deformation is performed in the austenite
recrystallisation region, austenite grains can be refined through recrystallisation. Then, when the
grains are further rolled in the austenite non-recrystallisation region, the refined austenite grains
become pancake shaped with dislocations inside. As a result, more abundant nucleation sites for
phase transformation could be made under condition ② , inducing a more effective grain
refinement compared to rolling condition ③ and other rolling conditions [546]. As coarse
precipitates commonly appear when rolling is processed in the austenite non-recrystallisation
region [22], rolling condition ③ would generate more coarse precipitates compared to rolling
condition ② because most deformation in rolling condition ③ is performed in the austenite
non-recrystallisation region. It should be noted that even though rolling condition ② can
achieve the highest YS of 907 MPa, UTS of 965 MPa and elongation of 25.8% among all the
rolling conditions, the impact properties of the processed steel is quite low, which limits the use
for structural applications. Therefore, rolling condition ③ is regarded as the optimal schedule. It
can generate YS of 860 MPa, UTS of 951 MPa and elongation of 23.5%. The contributions of
grain refinement and precipitation strengthening to the YS are 320 and 276 MPa, respectively,
under rolling condition ③ [545].

Fig. 64. Schematic illustration of TMP schedules with five rolling conditions [545]. SRT:
start rolling temperature, FRT: finish rolling temperature.

Rolling in the regions through austenite recrystallisation to austenite non-recrystallisation is


thought to be effective to maximise the use of grain refining and precipitation strengthening in
Nano Hiten steels, and a large number of researchers have applied this technology to the steels’
TMP schedules [525,537,539-541]. Recently, Ning et al. [542] proposed a new method of using
DIFT (route B) in the TMP of Nano Hiten steels, and made a comparison with the conventional
rolling processed in the austenite recrystallisation and non-recrystallisation regions (route A), as

96
indicated in Fig. 65. Their results indicated that the DIFT-processed sample had much finer
grain size compared to the sample processed by route A. Also, only fine precipitates (< 10 nm)
were observed in the DIFT-processed sample, and 93.6% of them were smaller than 5 nm. In
contrast, the sample processed by route B contained a number of coarse prcipitates with a
diameter larger than 10 nm. In DIFT, the phase transformation driving force is increased due to
the addition of deformation stored energy, which leads to an increase in the Ae3 temperature. As
a result, the phase transformation would be greatly accelerated when deformation is processed at
a temperature between Ae3 and Ar3 [547]. Compared to the sample processed by route A, the
degree of recovery in the DIFT-processed sample is quite low. A mass of dislocations
accumulated during the DIFT process would greatly increase the number of nucleation sites of
ferrite, which induces refinement of grains after austenite-to-ferrite transformation. Because
deformation and phase transformation occur almost simultaneously during DIFT rolling,
deformation can be well accommodated, and the existing time of dislocations in austenite is
significantly reduced. Concequently, strain-induced precipitation can be prevented. Therefore,
much more finer precipitates readily form in the DIFT-processed sample. DIFT provides an
alternative way of processing Nano Hiten steels with the purpose of refining both ferrite grains
and precipitates. More work on the optimisation of TMP schedules of Nano Hiten steels
considering rolling and the following isothermal treatment strategies can be conducted in future.
It should be noted that the load of rolling performed in the DIFT region is high compared to
conventional rolling due to the relatively low rolling temperature. Rolling capacity of a rolling
mill should also be considered in the DIFT rolling of Nano Hiten steels.

Fig. 65. Schematic illustration of TMP schedules comparing DIFT rolling (route B) with rolling in
austenite non-recrystallisation region (route A) [542]. SRT: start rolling temperature, FRT: finish
rolling temperature.

For stage (iii) illustrated in Fig. 63, the cooling rate of the Nano Hiten steels should be high
enough to avoid austenite decomposition and induce a high degree of supersaturation in the

97
ferrite matrix. The supersatured solid-solution ferrite can produce plenty of nano-sized interface
precipitates at the following coiling process. In laboratory research, the cooling rate is usually
controlled at 20 ℃/s or higher [527,530]. In practical rolling process, steels are commonly
cooled either in the air or by water. The cooling rate in air is approximately 10 ℃/s [525,541],
while the cooling rate by water may vary from 18 to 50 ℃/s depending on the dimensions and
hardenability of the steels [538,540].
Isothermal treatment is a key process in controlling the formation of a single ferrite phase
and nano-sized prcipitates during austenite-to-ferrite transformation. It is generally accepted that
a lower isothermal temperature benefits the refining of both ferrite grains and precipitates, and
the decrease of the row spacing of interface precipitates, and thereby the mprovement of the
final mechanical properties of the processed steel [534]. At a low isothermal temperature, the
driving force increases as supercooling increases, and hence, the nucleation rates of ferrite and
precipitates increase [541]. Huang et al. [540] found that the average ferrite grain size decreased
from 6.2 to 3.5 μm, and the row spacing of interface precipitates decreased from 31.6 to 13.4
nm, when the isothermal temperature was decreased from 700 to 600 ℃. As a result, the UTS
and YS could be increased from 720 and 670 MPa to 995 and 945 MPa, respectively, when the
isothermal temperature was decreased from 700 to 600 ℃, while only a slight loss in elongation
from 24.62% to 20.18% was observed. The work of Kim et al. [22] showed that the maximum
precipitation strengthening could be achieved when the isothermal temperature was maintained
at 620 ℃, while either a high (670 ℃) or a low (570 ℃) isothermal temperature did not
contribute effectively to the precipitation strengthening. Zhang et al. [543] found that full ferrite
was formed at isothermal temperatures of 600 and 650 ℃, but bainitic ferrite appeared at lower
isothermal temperatures of 500 and 550 ℃. Phaniraj et al. [539] suggested using 650 ℃ as the
isothermal temperature because (Ti,Mo)C carbides are expected to form at this temperature.
However, the isothermal temperature of 600 ℃ was found to be better than 650 ℃ based on the
fact that the number density of nano-sized (Ti,Mo)C carbides was higher and the grain size of
ferrite was finer at 600 ℃, which contributed to the best combination of strength and ductility
[543]. Therefore, an optimal range of isothermal temperature should be used in the TMP of
Nano Hiten steels, so that refinement of both polygonal ferrite grains and preicipitates can be
achieved simultaneously to the maximum possible extent. In practice, bainitic transformation
temperature can be obtained at first, while the isothermal temperature should be set higher than
this temperature. Based on the research by Funakawa et al. [525], the Ar3 temperaure determined
by dilatometric experiments can be used as the isothermal temperature. This, however, is not
always the best approach, and more experimental work is still needed to determine the optimal
isothermal temperature in the TMP of Nano Hiten steels. Coarsening of precipitates would

98
become more serious with the increase of holding time at the isothermal temperature. The
isothermal holding time is commonly set at 1 h to allow complete transformation of austenite-
to-ferrite and precipitation of nano-sized interface precipitates, as well as achievement of an
optimal combination of refined ferrite grains and precipitates [22,525,540,541,543,545,546].
The slow cooling in stage (v) (Fig. 63) simulates the coiling process of Nano Hien steels, and
furnace cooling is widely used for this purpose.
The contributions of grain refinement strengthening and precipitation strengthening to the
YS of the TMP-processed Nano Hiten steels are key criteria to evaluate the appropriateness of
TMP technologies. Grain refinement strengthening (  g ) can be calculated by using the Hall-

Petch equation [548,549], and precipitation strengthening (  p ) can be obtained by the Ashby-

Orowan equation [550,551], as expressed by Eqs. (5) and (6), respectively.


1 / 2
 g  k y d f (5)

K 1/ 2 d p
 p  f p ln (6)
dp b

where k y and K are constants, d f is the average ferrite grain size, d p is the average diameter

of precipitates, f p is the volume fraction of precipitates, and b is the Burgers vector.

Table 4 summarises the contributions of  g and  p to the YS of some Ti-Mo bearing

Nano Hiten steels after TMP treatement. It is clear that the values of  g and  p are not

identical at the same isothermal temperature. Also, there is no a monotonic varying trend for
either  g or  p with the change of isothermal temperature. This is because the chemical

compositions of the steels and/or their processing conditions are different, which induces
differences in ferrite grain size, precipitate size and volume fraction of precipitates. Grain
refinement and precipitation strengthening are the two most important components contributing
to improving the strength of Nano Hiten steels. It is possible to maximise the improvement of
mechanical properties of Nano Hiten steels with a combined effect of ferrite grain refinement
and precipitation strengthening by using optimised TMP technologies. Currently, Nano Hiten
steels with YS of 1180 MPa are available [524], and many more new grades of Nano Hiten
steels with YS over 1 GPa are expected in the future, with the development of Nano Hiten steels
and their TMP technologies.

99
Table 4 Contributions of  g and  p to the YS of some Ti-Mo bearing Nano Hiten steels.
 g  p Start rolling temperature Finish rolling temperature Isothermal temperature References
(℃) (℃) (℃)
(MPa) (MPa)
372 109 1200 870 500 [543]
362 157 1200 870 550 [543]
318 218 980 880 570 [22]
194 271 1150 1050 570 [22]
362 324 1200 870 600 [543]
216 403 N/A 900 600 [552]
304 430 1150 900 600 [540]
352 182 850 750 620 [545]
311 154 900 800 620 [545]
320 276 980 880 620 [545]
~260 300 N/A 900 620 [525]
401 199 1060 960 620 [546]
211 320 1150 1050 620 [546]
265 357 1150 900 625 [540]
353 274 1200 870 650 [543]
205 333 N/A 900 650 [552]
318 214 980 880 670 [22]
208 270 1150 1050 670 [22]
197 275 N/A 900 700 [552]
210 246 1150 900 700 [540]
Notes: “N/A” means no relevant information is found in the literature.

6. Summary and outlook


AHSSs are complex and sophisticated materials, with carefully selected chemical
compositions and multiphase microstructures resulting from precisely controlled TMP
technologies. Various strengthening mechanisms through control of TMP routes are available to
achieve a range of mechanical properties of AHSSs. This article examines the progress made in
the area of TMP of AHSSs to date, focusing on the relationships between processing,
microstructure and mechanical properties.
The first generation AHSSs typically get their formability from ferrite microstructure, and
strength from martensite microstructure, with the addition of low temperature transformation
products, such as bainite and carbon-enriched retained austenite, to achieve a good combination
of high strength and acceptable ductility. The TMP of the first generation AHSSs aims to fully
explore the potential of these phases in the control of the final mechanical properties. The key to
the TMP of the first generation steels is to moderately refine the microstructure, in combination
with the control of microstructural constituents, phase morphology and precipitation, with the
purpose of achieving an optimal balance between strength and ductility. The second generation
AHSSs are fundamentally austenite-based, and exhibit an excellent combination of strength and
formability. However, TMP-controlled rolling of the second generation AHSSs is still
challenging, especially for L-IP and TWIP steels, due to the elevated alloying additions that are
required to stabilise austenite at room temperature. The third generation AHSSs possess a better
strength/ductility combination than that exhibited by the first generation AHSSs, and are more

100
cost effective than the second generation AHSSs, which contain high-level alloying elements.
The TMP of the third generation AHSSs is currently focused on heat treatment because the
microstructural characteristics needed for the steels are primarily achieved by controlling post-
deformation heat treatment conditions. Through appropriate TMP strategies—especially the
control of processing parameters in quenching, partitioning, annealing and tempering—
excellent mechanical properties of the third generation AHSSs can be readily obtained. A single
ferrite phase strengthened by nano-sized precipitates in Nano Hiten steels can be easily obtained
through TMP. However, achieving maximum strengthening by ferrite grain refinement and
precipitation is still difficult because an optimal TMP route is affected by a number of
parameters, including the start rolling temperature, finish rolling temperature, isothermal
treatment temperature and holding time. An understanding of austenite recrystallisation and
non-recrystallisation behaviour, austenite-to-ferrite transformation and interface precipitation is
essential in the design of optimal TMP strategies of Nano Hiten steels.
The TMP presented in this article represents the recent progress in the processing of the first,
second and third generations of AHSSs, as well as Nano Hiten steels. As the discussion on the
mechanism of TMP-induced microstructural evolution and mechanical properties variation is
based on specified AHSSs under certain TMP schedules, the mechanism may be different with
the change of chemical compositions and processing conditions. As each grade of AHSSs has
unique chemical compositions, TMP should first be developed for each of the AHSSs to gain
deep insights into the processing-microstructure-property relationships, then to maximise the
obtaining of the desired microstructure—and therefore the targeted mechanical properties. To
achieve this, a careful consideration on the steel’s chemical compositions and the processing
parameters, including deformation temperature, strain, strain rate, cooling rate, heat treatment
temperature and holding time, must be made in the design of TMP strategies.
Driven by ongoing demand for high-performance lightweight materials from the automotive
and other manufacturing industries, the development of the TMP of AHSSs will undoubtedly
continue for some time. In order to eventually apply the TMP to produce high-performance and
high-quality AHSS products in the future, TMP schedules should not be complex and
processing parameters should remain close to practical conditions, so that the developed TMP
can be applied to the production lines of the steel-making industry. Currently, most research
studies on TMP have focused on improving the mechanical properties of AHSSs by optimising
processing parameters. The relationship between processing, microstructure, mechanical
properties and workability should also be considered in TMP schedule designs in order to avoid
defects, such as cracking, in the production of AHSSs.

101
Acknowledgement
The authors would like to thank Dr. Madeleine Strong Cincotta from the University of
Wollongong for assisting in English language editing.

102
References
[1] Cheah LW. Cars on a Diet: The Materials and Energy Impacts of Passenger Vehicle
Weight Reductions in the U.S. Massachusetts Institute of Technology, USA; 2010.
[2] Tolouei R, Titheridge H. Vehicle mass as a determinant of fuel consumption and
secondary safety performance. Transp Res Part D 2009;14:385-99.
[3] Broughton J. The British index for comparing the accident records of car models. Accid
Anal Prev 1996;28:101-9.
[4] Demeri MY. Advanced High-Strength Steels: Science, Technology, and Applications.
ASM International; 2013.
[5] Rowe J. Advanced Materials in Automotive Engineering. Woodhead Publishing
Limited; 2012.
[6] Galán J, Samek L, Verleysen P, Verbeken K, Houbaert Y. Advanced high strength
steels for automotive industry. Revista De Metall 2012;48(2):118-31.
[7] U.S. Energy Information Administration. Annual Energy Outlook 2014 Early Release
Overview. 2014.
[8] World Steel Association. Environmental case study: An advanced high-strength steel
family car. 2008.
[9] Jin YS. Development of advanced high strength steels for automotive applications. La
Metall Italiana 2011;6:43-8.
[10] Matlock DK, Speer JG. Design considerations for the next generation of advanced high
strength sheet steels. In Lee HC (ed.). Proceeding of the 3rd International Conference
on Structural Steels. The Korean Institute of Metals and Materials, Korea; 2006. pp.
774-81.
[11] WordAutoSteel. Advanced high strength steel (AHSS) application guidelines (version
4.1). WorldAutoSteel; June 2009.
[12] Tomita K, Funakawa Y, Shiozaki T, Maeda E, Yamamoto T. Hot rolled high strength
steel sheet strengthened by nano size precipitates-development of 780 MPa grade Nano
Hiten. Materia Jpn 2003;42:70-72.
[13] Sekita T, Kaneto S, Hasuno S, Sato A, Ogawa T, Ogura K. Materials and technologies
for automotive use. JFE Tech Rep 2004;2:1-18.
[14] Shimizu T, Funakawa Y, Kaneko S. High strength steel sheets for automobile
suspension and chassis use-high strength hot-rolled steel sheets with excellent press
formability and durability for critical safety parts-. JFE Tech Rep 2004;4:25-31.
[15] Davies G. Materials for Automotive Bodies. Butterworth Heinemann; 2012.
[16] Verlinden B, Driver J, Samajdar I, Doherty RD. Thermo-Mechanical Processing of
Metallic Materials. Elsevier; 2007.
[17] DeGarmo EP, Black JT, Kohser RA. Materials and Process in Manufacturing (9th
edition). Wiley; 2003.
[18] Gong P, Palmiere EJ, Rainforth WM. Thermomechanical processing route to achieve
ultrafine grains in low carbon microalloyed steels. Acta Mater 2016;119:43-54.
[19] Gong P, Palmiere EJ, Rainforth WM. Dissolution and precipitation behaviour in steels
microalloyed with niobium during thermomechanical processing. Acta Mater
2015;97:392-403.
[20] Cao J, Yan J, Zhang J, Yu T. Effects of thermomechanical processing on microstructure
and properties of bainitic work hardening steel. Mater Sci Eng A 2015;639:192-7.
[21] Nezakat M, Akhiani H, Hoseini M, Szpunar J. Effect of thermo-mechanical processing
on texture evolution in austenitic stainless steel 316L. Mater Charact 2014;98:10-7.
[22] Kim YW, Song SW, Seo SJ, Hong SG, Lee CS. Development of Ti and Mo micro-
alloyed hot-rolled high strength sheet steel by controlling thermomechanical controlled
processing schedule. Mater Sci Eng A 2013;565:430-8.
[23] Kong X, Lan L, Hu Z, Li B, Sui T. Optimization of mechanical properties of high
strength bainitic steel using thermo-mechanical control and accelerated cooling process.
103
J Mater Process Technol 2015;217:202-10.
[24] Militzer M. Thermomechanical processed steels. In: Hashmi MSJ (ed.). Comprehensive
Materials Processing. Elsevier; 2014.
[25] M. Venkatraman, T. Venugopalan, Issues of divergence in hot strip rolling. Proceeding
of the 2nd International Conference on Thermomechanical Processing of Steels
(TMP’2004), Liege, Belgium; 15-17 June 2004. pp. 99-106.
[26] Yao CK, Zhang YM, Men XY, Zhang SQ. Transformation to pearlite from austenitized
and recrystallized austenite. Mater Sci Eng 1986;83:L1-L6.
[27] Tamura I, Sekine H, Tanaka T, Ouchi C. Thermomechanical Processing of High-
Strength Low-Alloy Steels. Butterworth & Co. (Publishers) Ltd; 1988.
[28] Zackay VF. Thermomechanical processing. Mater Sci Eng 1976;25:247-61.
[29] Poliak EI, Pottore NS, Skolly RM, Umlauf WP, Brannbacka JC. Thermomechanical
processing of advanced high strength steels in production hot strip rolling. La Metall
Ital 2009;2:1-8.
[30] Zhao J, Jiang Z. Rolling of Advanced High Strength Steels: Theory, Simulation and
Practice. CRC Press; 2017.
[31] Jha G, Das S, Sinha S, Lodh A, Haldar A. Design and development of precipitate
strengthened advanced high strength steel for automotive application. Mater Sci Eng
A 2013;561:394-402.
[32] Zhu K, Barbier D, Iung T. Characterization and quantification methods of complex
BCC matrix microstructures in advanced high strength steels. J Mater Sci 2013;48:413-
23.
[33] Wang XZ, Hasood SH. Investigation of die radius arc profile on wear behaviour in
sheet metal processing of advanced high strength steels. Mater Des 2011;32:1118-28.
[34] Naderi M, Abbasi M, Saeed-Akbari A. Enhanced mechanical properties of a hot-
stamped advanced high-strength steel via tempering treatment. Metall Mater Trans A
2013;44:1852-61.
[35] Maeno T, Mori KI, Nagai T. Improvement in formability by control of temperature
in hot stamping of ultra-high strength steel parts. CIRP Annals - Manuf Technol
2014;63:301-4.
[36] Zhong N, Wang XD, Wang L, Rong YH. Enhancement of the mechanical properties of
a Nb-microalloyed advanced high-strength steel treated by quenching-partitioning-
tempering process. Mater Sci Eng A 2009;506:111-6.
[37] Koyama M, Akiyama E, Lee YK, Raabe D, Tsuzaki K. Overview of hydrogen
embrittlement in high-Mn steels, Int J Hydrogen Engery 2017;42:12706-23.
[38] Liu Q, Zhou Q, Venezuela J, Zhang M, Atrens A. Hydrogen influcence on some
adanced high-strength steels. Corros Sci 2017;125:114-138.
[39] Venezuela J, Liu Q, Zhang M, Zhou Q, Atrens A. A review of hydrogen embrittlement
of martensitic advanced high-strength steels. Corros Rev 2016;34:153-86.
[40] Bae GH, Huh H. Comparison of the optimum designs of center pillar assembly of an
auto-body between conventional steel and AHSS with a simplified side impact analysis.
Int J Automot Technol 2012;13:205-13.
[41] Rana R, Singh SB. Automotive Steels: Design, Metallurgy, Processing and Applications.
Woodhead Publishing; 2016.
[42] WordAutoSteel. Advanced high-strength steels application guidelines (version 5.0).
WorldAutoSteel; May 2014.
[43] Maffei B, Salvatore W, Valentini R. Dual-phase steel rebars for high-ductile r.c.
structures, Part 1: Microstructural and mechanical characterization of steel rebars. Eng
Struct 2007;29:3325-32.
[44] Zhao J, Jiang Z, Lee CS. Functions of Tungsten Alloying in Microalloyed Steels. Nova
Science Publishers; 2014.
[45] Zhao J, Jiang Z, Lee CS. Effects of tungsten addition and heat treatment conditions on
microstructure and mechanical properties of microalloyed forging steels. Mater Sci Eng
104
A 2013;562:144-51.
[46] Tasan CC, Diehl M, Yan D, Bechtold M, Roters F, Schemmann L, Zheng C, Peranio N,
Ponge D, Koyama M, Tsuzaki K, Raabe D. An overview of dual-phase steels: advances
in microstructure-oriented processing and micromechanically guided design. Annu Rev
Mater Res 2015;45:391-431.
[47] Perzyński K, Wrożyna A, Kuziak R, Legwand A, Madej L. Development and validation
of multi scale failure model for dual phase steels. Finite Elem Anal Des 2017;124:7-21.
[48] Chang PH, Preban AG. The effect of ferrite grain size and martensite volume fraction
on the tensile properties of dual phase steel. Acta Metall 1985;33:897-903.
[49] Jena AK, Chaturvedi MC. On the effect of the volume fraction of martensite on the
tensile strength of dual-phase steel. Mater Sci Eng 1988;100:1-6.
[50] Lai Q, Brassart L, Bouaziz O, Gouné M, Verdier M, Parry G, Perlade A, Bréchet Y,
Pardoen T. Influence of martensite volume fraction and hardness on the plastic bahavior
of dual-phase steels: Experiments and micromechanical modeling. Int J Plast
2016;80:187-203.
[51] Pouranvari M. Tensile strength and ductility of ferrite-martensite dual phase steels.
Assoc Metall Eng Serbia 2010;16(3):187-94
[52] Sudersanan PD, Kori N, Aprameyan S, Kempaiah DUN. The effect of carbon content in
martensite on the strength of dual phase steel. Bonfring Int J Ind Eng Manage Sci
2012;2(2):1-4.
[53] Saikaly W, Bano X, Issartel C, Rigaut G, Charrin L, Charai A. The effects of
thermomechanical processing on the precipitation in an industrial dual-phase steel
microalloyed with titanium. Metall Mater Trans A 2001;32:1939-47.
[54] Calcagnotto M, Adachi Y, Ponge D, Raabe D. Deformation and fracture mechanisms in
fine- and ultrafine-grained ferrite/martensite dual-phase steels and the effect of aging.
Acta Mater 2011;59:658-70.
[55] Delincé M, Bréchet Y, Embury JD, Geers MGD, Jacques PJ, Pardoen T. Structure–
property optimization of ultrafine-grained dual-phase steels using a microstructure-
based strain hardening model. Acta Mater 2007;55:2337-50.
[56] Son YI, Lee YK, Park KT, Lee CS, Shin DH. Ultrafine grained ferrite–martensite dual
phase steels fabricated via equal channel angular pressing: Microstructure and tensile
properties. Acta Mater 2005;53:3125-34.
[57] Park KT, Han SY, Ahn BD, Shin DH, Lee YK, Um KK. Ultrafine grained dual phase
steel fabricated by equal channel angular pressing and subsequent intercritical annealing.
Scr Mater 2004;51:909-13.
[58] Das D, Chattopadhyay PP. Influence of martensite morphology on the work-hardening
behavior of high strength ferrite-martensite dual-phase steel. J Mater Sci 2009;44:2957-
65.
[59] Terao N, Cauwe B. Influence of additional elements (Mo, Nb, Ta and B) on the
mechanical properties of high-manganese dual-phase steels. J Mater Sci 1988;23:1769-
78.
[60] Shin DH, Kim BC, Park KT, Choo WY. Microstructural changes in equal channel
angular pressed low carbon steel by static annealing. Acta Mater 2000;48:3245-52.
[61] Tsuji N. Ways to manage both strength and ductility in nanostructured steels. In: Weng
Y, Dong H, Gan Y (eds.). Advanced Steels: The Recent Scenario in Steel Science and
Technology. Springer; 2011. pp. 119-29.
[62] Karmakar A, Sivaprasad S, Nath SK, Misra RDK, Chakrabarti D. Comparison between
different processing schedules for the development of ultrafine-grained dual-phase steel.
Metall Mater Trans A 2014;45:2466-79.
[63] Karmakar A, Misra RDK, Neogy S, Chakrabarti D. Development of ultrafine-grained
dual-phase steels: mechanism of grain refinement during intercritical deformation.
Metall Mater Trans A 2013;44:4106-18.
[64] Zhao J, Lee JH, Kim YW, Jiang Z, Lee CS. Enhancing mechanical properties of a low-
105
carbon microalloyed cast steel by controlled heat treatment. Mater Sci Eng A
2013;559:427-35.
[65] Patra S, Roy S, Kumar V, Haldar A, Chakrabarti D. Ferrite grain size distributions in
ultra-fine-grained high-strength low-alloy steel after controlled thermomechanical
deformation. Metall Mater Trans A 2011;42:2575-90.
[66] Hodgson PD, Shokouhi A, Beladi H. A descriptive model for the formation of ultrafine
grained steels. ISIJ Int 2008;48:1046-9.
[67] Ferrieira JL, Melo TMF, Bott IS, Santos DB, Rios PR. Influence of thermomechanical
parameters on the competition between dynamic recrystallization and dynamic strain
induced transformation in C-Mn and C-Mn-Nb steels deformed by hot torsion. ISIJ Int
2007;47:1638-46.
[68] Santos DB, Bruzszek RK, Rodrigues PCM, Pereloma EV. Formation of ultrafine ferrite
microstructure in warm rolled and annealed C-Mn steel. Mater Sci Eng A
2003;346:189-95.
[69] Salehi AR, Serajzadeh S, Taheri AK. A study on the microstructural changes in hot
rolling of dual-phase steels. J Mater Sci 2006;41:1917-25.
[70] Matsumura Y, Yada H. Evolution of ultrafine-grained ferrite in hot successive
deformation. Trans ISIJ 1987;27:492-8.
[71] Basabe VV, Jonas JJ. The ferrite transformation in hot deformed 0.036% Nb austenite
at temperatures above the Ae3. ISIJ Int 2010;50:1185-92.
[72] Karmakar A, Misra RDK, Neogy S, Chakrabarti D. Development of ultrafine-grained
dual-phase steels: Mechanism of grain refinement during intercritical deformation.
Metall Mater Trans A 2013;44:4106-18.
[73] Mazaheri Y, Kermanpur A, Najafizadeh A, Saeidi N. Effects of initial microstructure
and thermomechanical processing parameters on microstructures and mechanical
properties of ultrafine grained dual phase steels. Mater Sci Eng A 2014;612:54-62.
[74] Speich GR, Demarest VA, Miller RL. Formation of austenite during intercritical
annealing of dual-phase steels. Metall Trans A 1981;12:1419-28.
[75] Judd RR, Paxton HW. Kinetics of Austenite formation from a spherodized Ferrite-
Carbide Aggregate. Trans Metall Soc AIME 1968;242:206-15.
[76] Azizi-Alizamini H, Militzer M, Poole WJ. Formation of ultrafine grained dual phase
steels through rapid heating. ISIJ Int 2011;51:958-64.
[77] Qu J, Dabboussi W, Hassani F, Nemes J, Yue S. Effect of microstructure on static and
dynamic mechanical property of a dual phase steel studied by shear punch testing. ISIJ
Int 2005;45:1741-6.
[78] Movahed P, Kolahgar S, Marashi SPH, Pouranvari M, Parvin N. The effect of
intercritical heat treatment temperature on the tensile properties and work hardening
behavior of ferrite–martensite dual phase steel sheets. Mater Sci Eng A 2009;518:1-6.
[79] Sun S, Pugh M. Properties of thermomechanically processed dual-phase steels
containing fibrous martensite. Mater Sci Eng A 2002;335:298-308.
[80] Sodjit S, Uthaisangsuk V. Microstructure based prediction of strain hardening behavior
of dual phase steels. Mater Des 2012;41:370-9.
[81] Mazaheri Y, Kermanpur A, Najafizadeh A. A novel route for development of ultrahigh
strength dual phase steels. Mater Sci Eng A 2014;619:1-11.
[82] Kot RA, Morris JW (eds.). Structure and Properties of Dual-Phase Steels. AIME, New
York; 1979. pp. 145-82.
[83] Asadi M, De Cooman BC, Palkowsk H. Influence of martensite volume fraction and
cooling rate on the properties of thermomechanically processed dual phase steel. Mater
Sci Eng A 2012;538:42-52.
[84] Mukherjee K, Hazra SS, Militzer M. Grain refinement in dual-phase steels. Metall
Mater Trans A 2009;40:2145-59.
[85] Beladi H, Kelly GL, Shokouhi A, Hodgson PD. The evolution of ultrafine ferrite
formation through dynamic stain-induced transformation. Mater Sci Eng A
106
2004;371:343-52.
[86] Hurley PJ, Hodgson PD. Formation of ultra-fine ferrite in hot rolled strip: potential
mechanisms for grain refinement. Mater Sci Eng A 2001;302:206-14.
[87] Choi JK, Seo DH, Lee JS, Um KK, Choo WY. Formation of ultrafine ferrite by strain-
induced dynamic transformation in plain low carbon steel. ISIJ Int 2003;43:746-54.
[88] Hong SC, Lim SH, Lee KJ, Lee KS. Determination of dynamic ferrite transformation
during deformation in austenite. In: Zhu YT, Langdon TG, Mishra RS, Semiatin SL,
Saran MJ, Lowe TC (eds.). Ultrafine-Grained Materials II. TMS, Warrendale, PA; 2002.
pp. 267-274.
[89] Beladi H, Kelly GL, Shokouhi A, Hodgson PD. Effect of thermomechanical parameters
on the critical strain for ultrafine ferrite formation through hot torsion testing. Mater Sci
Eng A 2004;367:152-61.
[90] Hong SC, Lee KS. Influence of deformation induced ferrite transformation on grain
refinement of dual phase steel. Mater Sci Eng A 2002;323:148-59.
[91] Yang Z, Wang R. Formation of ultra-fine grain structure of plain low carbon steel
through deformation induced ferrite transformation. ISIJ Int 2003;43:761-6.
[92] Mazaheri Y, Kermanpur A, Najafizadeh A. Nanoindentation study of ferrite-martensite
dual phase steels developed by a new thermomechanical processing. Mater Sci Eng A
2015;639:8-14.
[93] Adamczyk J, Grajcar A. Structure and mechanical properties of DP-type and TRIP-type
sheets obtained after the thermomechanical processing. J Mater Process Technol
2005;162-163:267-74
[94] Calcagnotto M, Ponge D, Raabe D. On the effect of manganese on grain size stability
and hardenability in ultrafine-grained ferrite/martensite dual-phase steels. Metall Mater
Trans A 2012;43:37-46.
[95] Calcagnotto M, Ponge D, Raabe D. Microstructure control during fabrication of
ultrafine grained dual-phase steel: Characterization and effect of intercritical annealing
parameters. ISIJ Int 2012;52:874-83.
[96] Rao MP, Sarma VS, Sankaran S. Development of high strength and ductile ultra fine
grained dual phase steel with nano sized carbide precipitates in a V-Nb microalloyed
steel. Mater Sci Eng A 2013;568:171-5.
[97] Chakrabarti D, Davis C, Strangwood M. Development of bimodal grain structures in
Nb-containing high-strength low-alloy steels during slab reheating. Metall Mater Trans
A 2008;39:1963-77.
[98] Chakrabarti D, Davis CL, Strangwood M. Effect of deformation and Nb segregation on
grain size bimodality in HSLA steel. Mater Sci Technol 2009;25:939-46.
[99] Cai S, Boyd JD. Mechanism of microstructural banding in hot rolled microalloyed
steels. Mater Sci Forum 2005;500-501:171-8.
[100] Karmakar A, Karani A, Patra S, Chakrabarti D. Development of bimodal ferrite-grain
structures in low-carbon steel using rapid intercritical annealing. Metall Mater Trans A
2013;44:2041-52.
[101] Azizi-Alizamini H, Militzer M, Poole WJ. A novel technique for developing bimodal
grain size distribution in low carbon steels. Scr Mater 2007;57:1065-8.
[102] Wang TS, Li Z, Zhang B, Zhang XJ, Deng JM, Zhang FC. High tensile ductility and
high strength in ultrafine-grained low-carbon steel. Mater Sci Eng A 2010;527:2798-
801.
[103] Rao MP, Sarma VS, Sankaran S. Processing of bimodal grain-sized ultrafine-grained
dual phase microalloyed V-Nb steel with 1370 MPa strength and 16 pct uniform
elongation through warm rolling and intercritical annealing. Metall Mater Trans A
2014;45:5313-7.
[104] Wang GD. New generation TMCP and innovative hot rolling process. J Northeast Univ
(Nat Sci) 2009;30:913-22.
[105] Liu EY, Peng LG, Yuan G, Wang ZD, Zhang DH, Wang GD. Advanced run-out
107
cooling technology based on ultra fast cooling and laminar cooling in hot strip mill. J
Cent South Univ 2012;19:1341-5.
[106] Herman JC. Impact of new rolling and cooling technologies on tehrmomechanically
processed steels. Ironmak Steelmak 2001;28:159-63.
[107] Lucas A, Simon P, Bourdon G, Herman JC, Riche P, Neutjens J, Harlet P. Metallurgical
aspects of ultra fast cooling in front of the down-coiler. Steel Res Int 2004;75:139-46.
[108] Bhattacharya P, Samanta AN, Chakraborty S. Spray evaporative cooling to achieve
ultra fast cooling in runout table. Int J Therm Sci 2009;48:1741-7.
[109] Cai X, Liu C, Liu Z. Process design and prediction of mechanical properties of dual
phase steels with prepositional ultra fast cooling. Mater Des 2014;53:998-1004.
[110] Tomida T, Imai N, Miyata K, Fukushima S, Yoshida M, Wakita M, Etou M, Sasaki T,
Haraguchi Y, Okada Y. Grain refinement of C-Mn steel to 1 μm by rapid cooling and
short interval multi-pass hot rolling in stable austenite region. ISIJ Int 2008;48:1148-57.
[111] Chiang J, Boyd JD, Pilkey AK. Effect of microstructure on retained austenite stability
and tensile behaviour in an aluminum-alloyed TRIP steel. Mater Sci Eng A
2015;638:132-42.
[112] Yi HL. Review on δ-transformation-induced plasticity (TRIP) steels with low density:
The concept and current progress. JOM 2014;66:1759-69.
[113] Li L, Wollants P, He YL, De Cooman BC, Wei XC, Xu ZY. Review and prospect of
high strength low alloy TRIP steel. Acta Metall Sin (Engl Lett) 2003;16:457-65.
[114] Jacques PJ, Girault E, Harlet P, Delannay F. The developments of cold-rolled TRIP-
assisted multiphase steels. Low silicon TRIP-assisted multiphase steels. ISIJ Int
2001;41:1061-7.
[115] Lee CG, Kim SJ, Lee TH, Lee S. Effects of volume fraction and stability of retained
austenite on formability in 0.1C-1.5Si-1.5Mn-0.5Cu TRIP-aided cold-rolled steel sheet.
Mater Sci Eng A 2004;371:16-23.
[116] De Mayer M, Vanderschueren D, De Cooman BC. The influence of the substitution of
Si by Al on the properties of cold rolled C-Mn-Si TRIP steels. ISIJ Int 1999;39:813-22.
[117] Pereloma EV, Timokhina IB, Miller MK, Hodgson PD. Three-dimensional atom probe
analysis of solute distribution in thermomechanically processed TRIP steels. Acta Mater
2007;55:2587-98.
[118] Tommita Y. Effect of bainitic transformation on mechanical properties of 0.6C-Si-Mn
steel. J Mater Sci 1995;30:105-10.
[119] Wang J, Zwaag SVD. Stabilization mechanisms of retained austenite in transformation-
induced plasticity steel. Metall Mater Trans A 2001;32:1527-39.
[120] Matsumura O, Sakuma Y, Takechi H. Enhancement of elongation by retained austenite
in intercritical annealed 0.4C-1.5Si-0.8Mn steel. Trans ISIJ 1987;27:570-9.
[121] Blonde R, Jimenez-Melero E, Zhao L, Wright JP, Bruck E, van der Zwaag S, van Dijk
NH. High-energy X-ray diffraction study on the temperature-dependent mechanical
stability of retained austenite in low-alloyed TRIP steels. Acta Mater 2012;60:565-77.
[122] Kuziak R, Kawalla R, Waengler S. Advanced high strength steels for automotive
industry. Arch Civ Mech Eng 2008;VIII(2):103-17.
[123] Mintz B. Hot dip galvanising of transformation induced plasticity and other
intercritically annealed steels. Int Mater Rev 2001;46:169-97.
[124] Bhadeshia HKDH, Edmonds DV. The bainite transformation in a silicon steel. Metall
Trans A 1989;10:895-907.
[125] Sugimoto K, Osui N, Kobayashi M, Hashimoto S. Plastic stability of retained austenite
in the cold-rolled 0.14%C-1.9%Si-1.7%Mn sheet steel. ISIJ Int 1995;35:1121-7.
[126] Hosseini SMK, Zarei-Hanzaki A, Yue S. Effects of ferrite phase characteristics on
microstructure and mechanical properties of thermomechanically-processed low-silicon
content TRIP-assisted steels. Mater Sci Eng A 2015;626:229-36.
[127] Barbé L, Verbeken K, Wettinck E. Effect of the addition of P on the mechanical
properties of low alloyed TRIP steels. ISIJ Int 2006;46:1251-7.
108
[128] Lim N, Park H, Kim S, Park C. Effects of aluminium on the microstructure and phase
transformation of TRIP steels. Met Mater Int 2012;18:647-54.
[129] Jacques PJ, Girault E, Mertens A, Verlinden B, Van Humbeeck J, Delannay F. The
developments of cold-rolled TRIP-assisted multiphase steels. Al-alloyed TRIP-assisted
multiphase steels. ISIJ Int 2001;41:1068-74.
[130] Bellhouse E, Mertens A, McDermid J. Development of the surface structure of TRIP
steels prior to hot-dip galvanizing. Mater Sci Eng A 2007;463:147-56.
[131] Bellhouse EM, McDermid JR. Analysis of the Fe–Zn interface of galvanized high Al–
low Si TRIP steels. Mater Sci Eng A 2008;491:39-46.
[132] Ishida K, Nishizawa T. Effect of alloying elements on stability of epsilon iron. Trans
Jpn Inst Met 1974;15:225-31.
[133] Jacques PJ, Girault E, Mertens A, Verlinden B, Van Humbeeck J, Delannay F. The
developments of cold-rolled TRIP-assisted multiphase steels. Al-alloyed TRIP-assisted
multiphase steels. ISIJ Int 2001;41:1068-74.
[134] Grajcar A, Zalecki W, Kuziak R. Designing of cooling conditions for Si-Al
microalloyed TRIP steel on the basis of DCCT diagrams. J Achievements Mater Manuf
Eng 2011;45(2):115-24.
[135] Mahieu J, Mak J, De Cooman BC, Claessens S. Phase transformation and mechanical
properties of Si-free CMnAl transformation-induced plasticity-aided steel. Metall Mater
Trans A 2002;33:2573-80.
[136] Wang C, Ding H, Cai M, Rolfe B. Multi-phase microstructure design of a novel high
strength TRIP steel through experimental methodology. Mater Sci Eng A
2014;610:436-44.
[137] Sudo M, Hashimoto S, Kambe S. Niobium bearing ferrite-bainite high strength hot-
rolled sheet steel with improved formability. Trans ISIJ 1983;23:303-11.
[138] Sakuma Y, Matlock DK, Krauss G. Intercritically annealed and isothermally
transformed 0.15 Pct C steels containing 1.2 Pct Si-1.5 Pct Mn and 4 Pct Ni: Part I.
transformation, microstructure, and room-temperature mechanical properties. Metall
Trans A 1992;23:1221-32.
[139] Girault E, Mertens E, Jacques P, Vanhumbeek J, Verlinden B, Delannay F. Comparison
of the effects of silicon and aluminium on the tensile behaviour of multiphase TRIP-
assisted steels. Scr Mater 2001;44:885-92.
[140] Chen HC, Era H, Shimizu M. Effect of phosphorus on the formation of retained
austenite and mechanical properties in Si-containing low-carbon steel sheet. Metall
Trans A 1989;20:437-45.
[141] Rigsbee JM. Proceedings of the ICOMAT’79. Cambridge, MA, USA; 1979. p. 381.
[142] Takahashi M, Bhadeshia HKDH. A model for the microstructure of some advanced
bainitic steels. Mater Trans JIM 1991;32:689-96.
[143] Li LF, Zhang XJ, Yang WY, Sun ZQ. Microstructure and mechanical properties of a
low-carbon Mn-Si multiphase steel based on dynamic transformation of undercooled
austenite. Metall Mater Trans A 2013;44:4337-45.
[144] Hashimoto S, Ikeda S, Sugimoto KI, Miyake S. Effects of Nb and Mo addition to
0.2%C-1.5%Si1.5%Mn steel on mechanical properties of hot rolled TRIP-aided steel
sheets. ISIJ Int 2004;44:1590-8.
[145] Pereloma EV, Timokhina IB, Hodgson PD. Transformation behaviour in
thermomechanically processed C–Mn–Si TRIP steels with and without Nb. Mater Sci
Eng A 1999;273-275:448-52.
[146] Fu B, Yang WY, Li LF, Sun ZQ. Effect of bainitic transformation temperature on the
mechanical behavior of cold-rolled TRIP steels studied with in-situ high-energy X-ray
diffraction. Mater Sci Eng A 2014;603:134-40.
[147] Zarei-Hanzaki A. Transformation Characteristics of Si-Mn TRIP Steels after
Thermomechanical Processing. McGill University, Canada; 1994.
[148] Timokhina IB, Pereloma EV, Hodgson PD. Microstructure and mechanical properties of
109
C-Si-Mn(-Nb) TRIP steels after simulated thermomechanical processing. Mater Sci
Technol 2001;17:135-40.
[149] Timokhina IB, Hodgson PD, Pereloma EV. Effect of microstructure on the stability of
retained austenite in transformation-induced-plasticity steels. Metall Mater Trans A
2004;35:2331-41.
[150] Yue S, DiChiro A, Zarei-Hanzaki A. Thermomechanical processing effects on C-Mn-Si
TRIP steels. JOM 1997;49:59-61.
[151] Shipway PH, Bhadeshia HKDH. Mechanical stabilization of bainite. Mater Sci Technol
1995;11:1116-28.
[152] Zarei-hanzaki A, Yue S. Ferrite formation characteristics in Si-Mn TRIP steels. ISIJ Int
1997;37:583-9.
[153] Li L, Wollants P, He YL, De Cooman BC, Wei XC, Xu ZY. Review and prospect of
high strength low alloy TRIP steel. Acta Metall Sin (Engl Lett) 2003;16:457-65.
[154] Hosseini SMK, Zaeri-Hanzaki A, Yue S. Effect of austenite deformation in non-
recrystallization region on microstructure development in low-silicon content TRIP-
assisted steels. Mater Sci Eng A 2014;618:63-70.
[155] Hosseini SMK, Zarei-Hanzaki A, Essadiqi E, Yue S. Effect of prior austenite
characteristics on mechanical properties of thermomechanically processed multiphase
TRIP steels. Mater Sci Technol 2008;24:1354-61.
[156] Umemoto M, Ohtsuka H, Tamura I. Transformation to pearlite from work-hardened
austenite. Trans ISIJ 1983;23:775-84.
[157] Skalova L, Divisova R, Jandova D. Thermo-mechanical processing of low-alloy TRIP-
steel. J Mater Process Technol 2006;175:387-92.
[158] Grajcar A. Hot-working in the γ+α region of TRIP-aided microalloyed steel. Arch
Mater Sci Eng 2007;28:743-50.
[159] Basuki A, Aernoudt E. Influence of rolling of TRIP steel in the intercritical region on
the stability of retained austenite. J Mater Process Technol 1999;89-90:37-43.
[160] Zareai-Hanzaki A, Hodgson PD, Yue S. The influcend of bainite on retained austenite
characteristics in Si-Mn TRIP steels. ISIJ Int 1995;35:79-85.
[161] Zrnik J, Stejskal O, Novy Z, Hornak P, Fujda M. Structure dependence of the TRIP
phenomenon in Si-Mn bulk steel. Mater Sci Eng A 2007;462:253-8.
[162] Zrnik J, Muransky O, Stejskal O, Lukas P, Hornak P. Effect of processing conditions on
structure development and mechanical response of Si-Mn ‘TRIP’ steel. Mater Sci Eng
A 2008;483-484:71-5.
[163] Muransky O, Sittner P, Zrnik J, Oliver EC. In situ neutron diffraction investigation of
the collaborative deformation-transformation mechanism in TRIP-assisted steels at
room and elevated temperatures. Acta Mater 2008;56:3367-79.
[164] Chen M, Wu R, Liu H, Wang L, Shi J, Dong H, Jin X. An ultrahigh strength steel
produced through deformation induced ferrite transformation and Q&P process. Sci
China: Technol Sci 2012;55:1827-32.
[165] Wakita M, Adachi Y, Tomata Y. Crystallography and mechanical properties of ultrafine
TRIP-aided multi-phase steels. Mater Sci Forum 2007;539-543:4351-6.
[166] Sun ZQ, Yang WY, Qi JJ, Hu AM. Deformation enhanced transformation and dynamic
recrystallization of ferrite in a low carbon steel during multipass hot deformation. Mater
Sci Eng A 2002;334:201-6.
[167] Yang WY, Qi JJ, Sun ZQ, Yang P. Characteristics of deformation enhanced
transformation in low carbon steel. Acta Metall Sin 2004;40:135-40.
[168] Yang WY, Li LF, Yin YY, Sun ZQ, Wang XT. Hot-rolled TRIP steels based on
dynamic transformation of undercooled austenite. Mater Sci Forum 2010;654-656:250-
3.
[169] Feng Q, Li L, Yang W, Sun Z. Effect of Nb on the stability of retained austenite in hot-
rolled TRIP steels based on dynamic transformation. Mater Sci Eng A 2014;603:169-75.
[170] Feng Q, Li L, Yang W, Sun Z. Microstructures and mechanical properties of hot-rolled
110
Nb-microalloyed TRIP steels by different thermos-mechanical processes. Mater Sci Eng
A 2014;605:14-21.
[171] Ranjan R, Beladi H, Singh SB, Hodgson PD. Thermo-mechanical processing of TRIP-
aided steels. Metall Mater Trans A 2015;46:3232-47.
[172] Guo Z, Li L. Influences of alloying elements on warm deformation behavior of high-Mn
TRIP steel with martensitic structure. Mater Des 2016;89:665-75.
[173] Guo Z, Li L, Yang W, Sun Z. Microstructures and mechanical properties of high-Mn
TRIP steel based on warm deformation of martensite. Metall Mater Trans A
2015;46:1704-14.
[174] De Moor E, Gibbs PJ, Speer JG, Matlock DK, Schroth JG. Strategies for third-
generation advanced high-strength steel development. AIST Trans 2010;7:132-44.
[175] Gibbs PJ, De Moor E, Merwin MJ, Clausen B, Speer JG, Matlock DK. Austenite
stability effects on tensile behavior of manganese-enriched-austenite transformation-
induced plasticity steel. Metall Mater Trans A 2011;42:3691-702.
[176] Arlazarov A, Goune M, Bouaziz O, Hazotte A, Petitgand G, Barges P. Evolution of
microstructure and mechanical properties of medium Mn steels during double annealing.
Mater Sci Eng A 2012;542:31-9.
[177] Puype A. Developing of advanced high strength steel via ultrafast annealing. Ghen
University, Belgium; 2014.
[178] Papaefthymiou S. A new opportunity for the design of advanced high strength steels
with heterogeneous-phase microstructures via rapid thermal processing. J Nanoscience
Adv Technol 2017;2(1):20-3
[179] Zhao J, Lee T, Lee JH, Jiang Z, Lee CS. Effects of tungsten addition on the
microstructure and mechanical properties of microalloyed forging steels. Metall Mater
Trans A 2013;44:3511-23.
[180] Hairer F, Karelova A, Krempaszky C, Werner E, Hebesberger T, Pichler A. Influence of
heat treatment on the microstructure and hardness of a low alloyed complex phase steel.
MS&T 2009: Proceedings from the Materials Science & Technology Conference.
MS&T Partner Societies, Pittsburgh, PA; 25-29 October 2009. pp. 78-83.
[181] Grajcar A. Thermodynamic analysis of precipitation processes in Nb-Ti-microalloyed
Si-Al TRIP steel. J Therm Anal Calorimetry 2014;118:1011-20.
[182] Adrian H. Thermodynamic model for precipitation of carbonitrides in high strength low
alloy steels containing up to three microalloying elements with or without additions of
aluminium. Mater Sci Technol 1992;8:406-20.
[183] Song SG, Kang KB, Park CG. Strain-induced precipitation of NbC in Nb and Nb-Ti
microalloyed HSLA steels. Scr Mater 2002;46:163-8.
[184] Wang K. A Study of HSLA Steels Microalloyed with Vanadium and Titanium during
Simulated Controlled Rolling Cycles. University of Canterbury, New Zealand; 2003.
[185] Lagneborg R, Siwecki T, Zajac S, Hutchinson B. The role of vanadium in microalloyed
steels. Scand J Metall 1999;28(5):186-241.
[186] Hong SG, Jun HJ, Kang KB, Park CG. Evolution of precipitates in the Nb-Ti-V
microalloyed HSLA steels during reheating. Scr Mater 2003;48:1201-6.
[187] Jung JG, Park JS, Kim J, Lee YK. Carbide precipitation kinetics in austenite of a Nb-Ti-
V microalloyed steel. Mater Sci Eng A 2011;528:5529-35.
[188] Timokhina IB, Enomoto M, Miller MK, Pereloma EV. Microstructure-property
relationship in the themomechanically processed C-Mn-Si-Nb-Al-(Mo) transformation-
induced plasticity steels before and after prestraining and bake hardening treatment.
Metall Mater Trans A 2012;43:2473-83.
[189] Fu B, Yang WY, Lu MY, Feng Q, Li LF, Sun ZQ. Microstructure and mechanical
properties of C-Mn-Al-Si hot-rolled TRIP steels with and without Nb based on dynamic
transformation. Mater Sci Eng A 2012;536:265-8.
[190] Palmiere EJ, Garcia CI, De Ardo AJ. Compositional and microstructural changes which
attend reheating and grain coarsening in steels containing niobium. Metall Mater Trans
111
A 1994;25:277-86.
[191] Wu H, Ju B, Tang D, Hu R, Guo A, Kang Q, Wang D. Effect Nb addition on the
microstructure and mechanical properties of an 1800 MPa ultrahigh strength steel. Mater
Sci Eng A 2015;622:61-6.
[192] Yamazaki K, Mizuyama Y, Oka M, Tsuchiya H, Yasuda H. Recent advances in
ultrahigh-strength sheet steels for automotive structural use. Nippon Steel Tech Rep
1995;64:37-44.
[193] Morris Jr JW, Kinney C, Pytlewski K, Adachi Y. Microstructure and cleavage in lath
martensitic steels. Sci Technol Adv Mater 2013;14:014208-1-014208-9.
[194] Mohrbacher H. Martensitic automotive steel sheet - Fundamentals and metallurgical
optimization strategies. Adv Mater Res 2015;1063:130-42.
[195] Galindo-Nava E, Rivera-Díaz-del-Castillo P. A model for the microstructure behaviour
and strength evolution in lath martensite. Acta Mater 2015;98:81-93.
[196] Matsuoka Y, Iwasaki T, Nakada N, Tsuchiyama T, Takaki S. Effect of grain size on
thermal and mechanical stability of austenite in metastable austenitic stainless steel. ISIJ
Int 2013;53:1224-30.
[197] Takaki S, Fukunaga K, Syarif J, Tsuchiyama T. Effect of grain refinement on thermal
stability of metastable austenitic steel. Mater Trans 2004;45:2245-51.
[198] Morito S, Saito H, Ogawa T, Furuhara T, Maki T. Effect of austenite grain size on the
morphology and crystallography of lath martensite in low carbon steels. ISIJ Int
2005;45:91-4.
[199] Hanamura T, Torizuka S, Tamura S, Enokida S, Takechi H. Effect of austenite grain
size on transformation behavior, microstructure and mechanical properties of 0.1C–5Mn
martensitic steel. ISIJ Int 2013;53:2218-25.
[200] Cui S, Cui Y, Wan J, Rong Y, Zhang J. Grain size dependence of the martensite
morphology – A phase-field study. Comput Mater Sci 2016;121:131-42.
[201] Morito S, Tanaka H, Konishi R, Furuhara T, Maki T. The morphology and
crystallography of lath martensite in Fe-C alloys. Acta Mater 2003;51:1789-99.
[202] Morito S, Huang X, Furuhara T, Maki T, Hansen N. The morphology and
crystallography of lath martensite in alloy steels. Acta Mater 2006;54:5323-31.
[203] Morris Jr JW. Comments on the microstructure and properties of ultrafine grained steel.
ISIJ Int 2008;48:1063-70.
[204] Swarr T, Krauss G. The effect of structure on the deformation of as-quenched and
tempered martensite in an Fe-0.2 pct C alloy. Metall Trans A 1976;7:41-8.
[205] Wang C, Wang M, Shi J, Hui W, Dong H. Effect of microstructure refinement on the
strength and toughness of low alloy martensitic steel. J Mater Sci Technol 2007;23:659-
64.
[206] Wang C, Wang M, Shi J, Hui W, Dong H. Effect of microstructural refinement on the
toughness of low carbon martensitic steel. Scr Mater 2008;58:492-5.
[207] Zhang C, Wang Q, Ren J, Li R, Wang M, Zhang F, Sun K. Effect of martensitic
morphology on mechanical properties of an as-quenched and tempered 25CrMo48V
steel. Mater Sci Eng A 2012;534:339-46.
[208] Bandyopadhyay PS, Ghosh SK, Kundu S, Chatterjee S. Evolution of microstructure and
mechanical properties of thermomechanically processed ultrahigh-strength steel. Metall
Mater Trans A 2011;42:2742-52.
[209] Bandyopadhyay PS, Kundu S, Ghosh SK, Chatterjee S. Structure and properties of a
low-carbon, microalloyed ultra-high-strength steel. Metall Mater Trans A
2011;42:1051-61.
[210] Kaijalainen AJ, Suikkanen P, Karjalainen LP, Jonas JJ. Effect of austenite pancaking on
the microstructure, texture and bendability of an ultrahigh-strength strip steel. Metall
Mater Trans A 2014;45:1273-83.
[211] Han Y, Shi J, Xu L, Cao WQ, Dong H. Effect of hot rolling temperature on grain size
and precipitation hardening in a Ti-microalloyed low-carbon martensitic steel. Mater
112
Sci Eng A 2012;553:192-9.
[212] Zhao J, Jiang Z, Lee CS. Enhancing impact fracture toughness and tensile properties of
a microalloyed cast steel by hot forging and post-forging heat treatment processes.
Mater Des 2013;47:227-33.
[213] Asadi M, Palkowski H. Influence of thermo-mechanical processing parameters and
chemical composition on bake hardening ability of hot rolled martensitic steels. Steel
Res Int 2009;80:499-506.
[214] Barbosa R, Boratto F, Yue S, Jonas JJ. The influence of chemical composition on the
recrystallization behavior of microalloyed steel. In: De Ardo AJ (ed.). THERMEC88
Conference Proceedings; 1988. pp. 51-61.
[215] Zhao J, Jiang Z, Kim JS, Lee CS. Effects of tungsten on continuous cooling
transformation characteristics of microalloyed steels. Mater Des 2013;49:252-8.
[216] Maalekian M, Kozeschnik E, Chatterjee S, Bhadeshia HKDH. Mechanical stabilisation
of eutectoid steel. Mater Sci Technol 2007;23:610-2.
[217] Wang HS, Yang JR, Bhadeshia HKDH. Characterisation of severely deformed
austenitic stainless steel wire. Mater Sci Technol 2005;21:1323-8.
[218] Maalekian M, Lendinez ML, Kozeschnik E, Brantner HP, Cerjak H. Effect of hot
plastic deformation of austenite on the transformation characteristics of eutectoid carbon
steel under fast heating and cooling conditions. Mater Sci Eng A 2007;454-455:446-52.
[219] Nikravesh M, Naderi M, Akbari GH. Influence of hot plastic deformation and cooling
rate on martensite and bainite start temperatures in 22MnB5 steel. Mater Sci Eng A
2012;540:24-9.
[220] Xiao FR, Liao B, Qiao GY, Guan SZ. Effect of hot deformation on phase
transformation kinetics of 86CrMoV7 steel. Mater Charact 2006;57:306-13.
[221] Reza T, Abbas N, Reza S. Drawing of CCCT diagrams by static deformation and
consideration deformation effect on martensite and bainite transformation in NiCrMoV
steel. J Mater Process Technol 2008;196:321-31.
[222] Wang HZ, Yang P, Mao WM, Lu FY. Effect of hot deformation of austenite on
martensitic transformation in high manganese steel. J Alloy Compd 2013;558:26-33.
[223] Lu F, Yang P, Meng L, Cui F, Ding H. Influences of thermal martensites and grain
orientations on strain-induced martensites in high manganese TRIP/TWIP steels. J
Mater Sci Technol 2011;27:257-65.
[224] Kim H, Park J, Kang M, Lee S. Interpretation of Charpy impact energy characteristics
by microstructural evolution of dynamically compressed specimens in three tempered
martensitic steels. Mater Sci Eng A 2016;649:57-67.
[225] Speich GR. Tempering of low-carbon martensite. Trans Metall Soc AIME
1969;245:2552-64.
[226] Speich GR, Leslie WC. Tempering of steel. Metall Trans 1972;3:1043-54.
[227] Wells MGH. An electron transmission study of the tempering of martensite in an Fe–Ni–
C alloy. Acta Metall 1964;12:389–99
[228] Caron RN, Krauss G. The tempering of Fe–C lath martensite. Metall Trans 1972;3:2381-9
[229] Nagakura S, Hirotsu Y, Kusunoki M, Suzuki T, Nakamuram Y. Crystallographic study of
the tempering of martensitic carbon steel by electron microscopy and diffraction. Metall
Trans A 1983;14:1025-31.
[230] Speich GR, Warlimont H. Yield strength and transformation substructure of low-carbon
martensite. J Iron Steel Inst 1968;206:385-92.
[231] Massardier V, Goune M, Fabregue D, Selouane A, Douillard T, Bouaziz O. Evolution of
microstructure and strength during the lutra-fast tempering of Fe-Mn-C martensitic steels.
J Mater Sci 2014;49:7782-96.
[232] Liu DY, Yang ZG, Bai BZ, Fang HS, Yang WY. The properties of 1500 MPa grade alloy
steel with carbide free bainite/martensite mixed microstructure. ISIJ Int 2003;43:433-7.
[233] Duan Z, Li Y, Zhang M, Shi M, Zhu F, Zhang S. Effects of quenching process on
mechanical properties and microstructure of high strength steel. J Wuhan Univ Tehchnol -
113
Mater Sci Ed 2012;27:1024-8.
[234] Prakash A, Hochrainer T, Reisacher E, Riedel H. Twinning models in self-consistent
texture simulations of TWIP steels. Steel Res Int 2008;79(8):645-52.
[235] Fischer FD, Schaden T, Appel F, Clemens H. Mechanical twins, their development and
growth. Eur J Mech A Solids 2003;22:709-26.
[236] Dumay A, Chateau JP, Allain S, Migot S, Bouaziz O. Influence of addition elements on
the stacking-fault energy and mechanical properties of an austenitic Fe–Mn–C steel.
Mater Sci Eng A 2008;483-484:184-7.
[237] Allain S, Chateau JP, Dahmoun D, Bouaziz O. Modeling of mechanical twinning in a
high manganese content austenitic steel. Mater Sci Eng A 2004;387-389:272-6.
[238] Jin JE, Lee YK. Strain hardening behavior of a Fe–18Mn–0.6C–1.5Al TWIP steel.
Mater Sci Eng A 2009;527:157-61.
[239] Gutierrez-Urrutia I, Raabe D. Dislocation and twin substructure evolution during strain
hardening of an Fe-22 wt.% Mn-0.6 wt.% C TWIP steel observed by electron
channeling contrast imaging. Acta Mater 2011;59:6449-62.
[240] Pierce DT, Jiménez, Bentley J, Raabe D, Oskay C, Wittig JE, The influence of
manganese content on the stacking fault and austenite/ε-martensite interfacial energies
in Fe-Mn-(Al-Si) steels investigated by experiment and theory. Acta Mater
2014;68:238-53.
[241] Hamada AS, Karjalainen LP, Somani MC. The influence of aluminum on hot
deformation behavior and tensile properties of high-Mn TWIP steels. Mater Sci Eng A
2007;467:114-24.
[242] Lehnhoff GR, Findley KO, De Cooman BC. The influence of silicon and aluminum
alloying on the lattic parameter and stacking energy of austenitic steel. Scr Mater
2014;92:19-22.
[243] Allain S, Chateau JP, Bouaziz O, Migot S, Guelton N. Correlations between the
calculated stacking fault energy and the plasticity mechanisms in Fe-Mn-C alloys.
Mater Sci Eng A 2004;387-389:158-62.
[244] Wan J, Chen S, Xu Z. The influence of temperature on stacking fault energy in Fe-
based alloys. Sci China Ser E 2001;44:345-52.
[245] Tiwari GP, Ramanujan RV. The relation between the electron to atom ratio and some
properties of metallic systems. J Mater Sci 2001;36:271-83.
[246] De Cooman BC, Chin KG, Kim J. High Mn TWIP steels for automotive applications. In:
Chiaberge M (ed.). New Trends and Developments in Automotive System Engineering.
InTech; 2011.
[247] Chen L, Zhao Y, Qin X. Some aspects of high manganese twinning-induced plasticity
(TWIP) steel, a review. Acta Metall Sin (Engl Lett) 2013;26:1-15.
[248] Lee JW, Wu CC, Liu TF. The influence of Cr alloying on microstructures of Fe–Al–
Mn–Cr alloys. Scr Mater 2004;50:1389-93.
[249] Tsay GD, Lin CL, Chao CG, Liu TF. A new austenitic FeMnAlCrC Alloy with high-
strength, high-ductility, and moderate corrosion resistance. Mater Trans 2010;51:2318-
21.
[250] Scott C, Remy B, Collet JL, Cael A, Bao C, Danoix F, Malard B, Curfs C. Precipitation
strengthening in high manganese austenitic TWIP steels. Int J Mater Res 2011;102:538-
49.
[251] Gutierrez-Urrutia I, Zaefferer S, Raabe D. The effect of grain size and grain orientation
on deformation twinning in a Fe-22 wt.% Mn-0.6 wt.% C TWIP steel. Mater Sci Eng A
2010;527:3552-60.
[252] Neu RW. Performance and characterization of TWIP steels for automotive applications.
Mater Perform Charact 2013;2:244-84.
[253] Tol RT, Kim JK, Zhao L, Sietsma J, De Cooman BC. α'-Martensite formation in deep-
drawn Mn-based TWIP steel. J Mater Sci 2012;47:4845-50.
[254] Beladi H, Timokhina IB, Estrin Y, Kim J, De Cooman BC, Kim SK. Orientation
114
dependence of twinning and strain hardening behaviour of a high manganese twinning
induced plasticity steel with polycrystalline structure. Acta Mater 2011;59:7787-99.
[255] El-Danaf E, Kalidindi SR, Doherty RD. Influence of grain size and staking-fault energy
on deformation twinning in fcc metals. Metall Mater Trans A 1999;30:1223-33.
[256] Venables JA. Deformation twinning in face-centred cubic metals. Philos Mag
1961;6:379-96.
[257] Fujita H, Mori T. A formation mechanism of mechanical twins in F.C.C. Metals. Scr
Metall 1975;9:631-6.
[258] Mahajan S, Chin GY. Formation of deformation twins in f.c.c. crystals. Acta Metall
1973;21:1353-63.
[259] Kwon O, Lee K, Kim G, Chin KG. New trends in advanced high strength steel
developments for automotive applications. Mater Sci Forum 2010;638-642:136-41.
[260] Bouaziz O, Allain S, Scott CP, Cugy P, Barbier D. High manganese austenitic twinning
induced plasticity steels: A review of the microstructure properties relationships. Curr
Opin Solid State Mater Sci 2011;15:141-68.
[261] De Cooman BC, Estrin Y, Kim SK. Twinning-induced plasticity (TWIP) steels. Acta
Mater 2018;142:283-362.
[262] Yen HW, Huang M, Scott CP, Yang JR. Interactions between deformation-induced
defects and carbides in a vanadium-containing TWIP steel. Scr Mater 2012;66:1018-23.
[263] Kusakin P, Belyakov A, Molodov DA, Kaibyshev R. On the effect of chemical
composition on yield strength of TWIP steels. Mater Sci Eng A 2017;687:82-4.
[264] Ghasri-Khouzani M, McDermid JR. Effect of carbon content on the mechanical
properties and microstructural evolution of Fe–22Mn–C steels. Mater Sci Eng A
2015;621:118-27.
[265] Jung IC, De Cooman BC. Temperature dependence of the flow stress of Fe-18Mn-0.6C-
xAl twinning-induced plasticity steel. Acta Mater 2013;61:6724-35.
[266] Jin JE, Lee YK. Effects of Al on microstructure and tensile properties of C-bearing high
Mn TWIP steel. Acta Mater 2012;60:1680-8.
[267] Hong S, Shin SY, Kim HS, Lee S, Kim SK, Chin KG, Kim NJ. Effects of aluminum
addition on tensile and cup forming properties of three twinning induced plasticity
steels. Metall Mater Trans A 2012;43:1870-83.
[268] Seol JB, Lim NS, Lee BH, Renaud L, Park CG. Atom probe tomography and nano
secondary ion mass spectroscopy investigation of the segregation of boron at austenite
grain boundaries in 0.5 wt.% carbon steels. Met Mater Int 2011;17:413-6.
[269] De Cooman BC. High Mn TWIP steel and medium Mn steel. In: Rana R, Singh SB
(eds.). Automotive Steels: Design, Metallurgy, Processing and Applications. Elsevier;
2017. pp. 317-85.
[270] Shun T, Wan CM, Byrne JG. A study of work hardening in austenitic Fe-Mn-C and Fe-
Mn-Al-C alloys. Acta Metall Mater 1992;40:3407-12.
[271] Hong S, Lee J, Lee BJ, Kim HS, Kim SK, Chin KG, Lee S. Effects of intergranular
carbide precipitation on delayed fracture behavior in three Twinning Induced Plasticity
(TWIP) steels. Mater Sci Eng A 2013;587:85-99.
[272] Oda K, Fujimura H, Ino H. Local interactions in carbon-carbon and carbon-M (M: Al,
Mn, Ni) atomic pairs in FCC γ-iron. J Phys Condens Matter 1994;6:679-92.
[273] Bartlett LN, Van Aken DC, Medvedeva J, Isheim D, Medvedeva NI, Song K. An atom
probe study of kappa carbide precipitation and the effect of silicon addition. Metall
Mater Trans A 2014;45:2421-35.
[274] Santos DB, Saleh AA, Gazder AA, Carman A, Duarte DM, Ribeiro EAS, Gonzalez BM,
Pereloma EV. Effect of annealing on the microstructure and mechanical properties of
cold rolled Fe–24Mn–3Al–2Si–1Ni–0.06C TWIP steel. Mater Scie Eng A
2011;528:3545-55.
[275] Kang S, Jung YS, Jun JH, Lee YK. Effects of recrystallization annealing temperature on
carbide precipitation, microstructure, and mechanical properties in Fe–18Mn–0.6C–
115
1.5Al TWIP steel. Mater Sci Eng A 2010;527:745-51.
[276] Bouaziz O, Scott CP, Petitgand G. Nanostructured steel with high work-hardening
by the exploitation of the thermal stability of mechanically induced twins. Scr
Mater 2009;60:714-6.
[277] Bracke L. Deformation Behaviour of Austenitic Fe-Mn alloys by Twinning and
Martensitic Transformation. University of Ghent, Belgium; 2007.
[278] Rahman KM, Vorontsov VA, Dye D. The effect of grain size on the twin initiation
stress in a TWIP steel. Acta Mater 2015;89:247-57.
[279] Yanushkevich Z, Belyakov A, Kaibyshev R, Haase C, Molodov DA. Effect of cold
rolling on recrystallization and tensile behavior of a high-Mn steel. Mater Charact
2016;112:180-7.
[280] Sabet M, Zarei-Hanzaki A, Khoddam S. Dynamic restoration processes in high-Mn
TWIP steels. J Eng Mater Technol 2009;131:044502-1-044502-5.
[281] Humphreys FJ, Hatherly M. Recrystallization and Related Annealing Phenomena (2nd
Edition). Elsevier; 2004.
[282] Khosravifard A, Hamada AS, Moshksar MM, Ebrahimi R, Porter DA, Karjalainen LP.
High temperature deformation behavior of two as-cast high-manganese TWIP steels.
Mater Sci Eng A 2013;582:15-21.
[283] Wang X, Brunger E, Gottstein G. The role of twinning during dynamic recrystallization in
alloy 800H. Scr Mater 2002;46:875-80.
[284] Field DP, Bradford LT. The Role of Annealing Twins During Recrystallization of Cu.
Acta Mater 2007;55:4233-41.
[285] Marcinkowski MJ, Miller DS. The effect of ordering on the strength and dislocation
arrangements in the Ni3Mn superlattice. Philoso Mag 1961;6:871-93.
[286] Copley SM, Kear BH. The dependence of the width of a dissociated dislocation on
dislocation velocity. Acta Metall 1968;16:227-31.
[287] Goodchild D, Roberts WT, Wilson DV. Plastic deformation and phase transformation in
textured austenitic stainless steel. Acta Metall 1970;18:1137-45.
[288] Karaman I, Sehitoglu H, Maier HJ, Chumlyakov YI. Competing mechanisms and
modeling of deformation in austenitic stainless steel single crystals with and without
nitrogen. Acta Mater 2001;49:3919-33.
[289] Eskandari M, Zarei-Hanzaki A, Szpunar JA, Mohtadi-Bonab MA, Kamali AR, Nazarian-
Samani M. Microstructure evolution and mechanical behavior of a new microalloyed high
Mn austenitic steel during compressive deformation. Mater Sci Eng A 2014;615:424-35.
[290] Hajkazemi J, Zarei-Hanzaki A, Sabet M, Khoddam S. Double-hit compression behavior
of TWIP steels. Mater Sci Eng A 2011;530:233-8.
[291] Grajcar A, Borek W. Thermo-mechanical processing of high-manganese austenitic TWIP-
type steels. Arch Civ Mech Eng 2008;8:29-38.
[292] Dobrzański LA, Grajcar A, Borek W. Microstructure evolution of high-manganese steel
during the thermomechanical processing. Arch Mater Sci Eng 2009;37:69-76.
[293] Dobrzański LA, Borek W. Hot deformation and recrystallization of advanced high-
manganese austenitic TWIP steels. J Achievements Mater Manuf Eng 2011;46:71-8.
[294] Dobrzański LA, Grajcar A, Borek W. Hot-working behaviour of high-manganese
austenitic steels. J Achievements Mater Manuf Eng 2008;31:7-14.
[295] Dobrzański LA, Grajcar A, Borek W. Influence of hot-working conditions on a
structure of high-manganese austenitic steels. J Achievements Mater Manuf Eng
2008;29:139-42.
[296] Dobrzański LA, Borek W. Thermo-mechanical treatment of Fe-Mn-(Al, Si)
TRIP/TWIP steels. Arch Civ Mech Eng 2012;12:299-304.
[297] Uranga P, Fernández AI, López B, Rodriguez-Ibabe JM. Transition between static and
metadynamic recrystallization kinetics in coarse Nb microalloyed austenite. Mater Sci
Eng A 2003;345:319-27.
[298] Doherty RD, Hughes DA, Humphreys FJ, Jonas JJ, Jensen DJ, Kassner ME, King WE,
116
McNelley TR, McQueen HJ, Rollett AD. Current issues in recrystallization: a review.
Mater Sci Eng A 1997;238:219-74.
[299] Kugler G, Turk R. Modeling the dynamic recrystallization under multi-stage hot
deformation. Acta Mater 2004;52:4659-68.
[300] Gomez M, Rancel L, Fernandez BJ, Medina SF. Evolution of austenite static
recrystallization and grain size during hot rolling of a V-microalloyed steel. Mater Sci
Eng A 2009;501:188-96.
[301] El-Danaf E, Kalidindi SR, Doherty RD. Influence of grain size and stacking-fault
energy on deformation twinning in fcc metals. Metall Mater Trans A 1999;30:1223-33.
[302] Kusakin PS, Kaibyshev RO. High-Mn twinning-induced plasticity steels: microstructure
and mechanical properties. Rev Adv Mater Sci 2016;44:326-60.
[303] Khalesian AR, Zarei-Hanzaki A, Abedi HR, Pilehva F. An investigation into the room
temperature mechanical properties and microstructural evolution of thermomechanically
processed TWIP steel. Mater Sci Eng A 2014;596:200-6.
[304] Mandal S, Bhaduri AK, Sarma VS. Influence of state of stress on dynamic
recrystallization in a titanium-modified austenitic stainless steel. Metall Mater Trans A
2012;43:410-4.
[305] Han Y, Liu G, Zou D, Liu R, Qiao G. Deformation behavior and microstructural
evolution of as-cast 904L austenitic stainless steel during hot compression. Mater Sci
Eng A 2013;565:342-50.
[306] Marandi A, Zarei-Hanzaki A, Haghdadi N, Eskandari M. The prediction of hot
deformation behavior in Fe–21Mn–2.5Si–1.5Al transformation-twinning induced
plasticity steel. Mater Sci Eng A 2012;554:72-8.
[307] Mandal S, Bhaduri AK, Sarma VS. Studies on twinning and grain boundary character
distribution during anomalous grain growth in a Ti-modified austenitic stainless steel.
Mater Sci Eng A 2009;515:134-40.
[308] Belyakov A, Kaibyshev R, Torganchuk V. Microstructure and mechanical properties of
18%Mn TWIP/TRIP steels processed by warm or hot rolling. Steel Res Int
2016;88(2):1600123.
[309] Sakai T, Belyakov A, Kaibyshev R, Miura H, Jonas JJ. Dynamic and post-dynamic
recrystallization under hot, cold and severe plastic deformation conditions. Prog Mater Sci
2014;60:130-207.
[310] Lee T, Koyama M, Tsuzaki K, Lee YH, Lee CS. Tensile deformation behavior of Fe–
Mn–C TWIP steel with ultrafine elongated grain structure. Mater Lett 2012;75:169-71.
[311] Koyama M, Sawaguchi T, Lee T, Lee CS, Tsuzaki K. Work hardening associated with
epsilon-martensitic transformation, deformation twinning and dynamic strain aging in Fe–
17Mn–0.6C and Fe–17Mn–0.8C TWIP steels. Mater Sci Eng A 2011;528:7310-6.
[312] Ueji R, Tsuchida N, Terada D, Tsuji N, Tanaka Y, Takemura A, Kunishige K. Tensile
properties and twinning behavior of high manganese austenitic steel with fine-grained
structure. Scr Mater 2008;59:963-6.
[313] Koyama M, Sawaguchi T, Tsuzaki K. Work hardening and uniform elongation of an
ultrafine-grained Fe–33Mn binary alloy. Mater Sci Eng A 2011;530:659-63.
[314] Lee S, Kim J, Lee SJ, De Cooman BC. Effect of nitrogen on the critical strain for dynamic
strain aging in high-manganese twinning-induced plasticity steel. Scr Mater 2011;65:528-
31.
[315] Zhang L, Liu XH, Shu KY. Microstructure and mechanical properties of hot-rolled Fe-
Mn-C-Si TWIP steel. J Iron Steel Res Int 2011;18(12):45-8.
[316] Wang SH, Liu ZY, Zhang WN, Wang GD, Liu JL, Liang GF. Microstructure and
mechanical property of strip in Fe-23Mn-3Si-3Al TWIP steel by twin roll casting. ISIJ
Int 2009;49:1340-6.
[317] McGuire MF. Stainless Steels for Design Engineers. ASM International; 2008.
[318] Hamada AS, Kisko AP, Sahu P, Karjalainen LP. Enhancement of mechanical properties
of a TRIP-aided austenitic stainless steel by controlled reversion annealing. Mater Sci
117
Eng A 2015;628:154-9.
[319] Behjati P, Kermanpur A, Najafizadeh A, Samaei Baghbadorani H. Effect of annealing
temperature on nano/ultrafine grain of Ni-free austenitic stainless steel. Mater Sci Eng
A 2014;592:77-82.
[320] Talha M, Kumar S, Behera CK, Sinha OP. Effect of cold working on biocompatibility
of Ni-free high nitrogen austenitic stainless steels using Dalton’s Lymphoma cell line.
Mater Sci Eng C 2014;35:77-84.
[321] Moon J, Ha HY, Lee TH. Corrosion behavior in high heat input welded heat-affected
zone of Ni-free high-nitrogen Fe-18Cr-10Mn-N austenitic stainless steel. Mater Charact
2013;82:113-9.
[322] Sugimoto K, Sawada Y. The role of molybdenum additions to austenitic stainless steels
in the inhibition of pitting in acid chloride solutions. Corros Sci 1977;17:425-45.
[323] Cunat PJ. Alloying elements in stainless steel and other chromium-containing alloys.
Euro Inox; 2004.
[324] Ahmedabadi P, Kain V, Arora K, Samajdar I, Sharma SC, Ravindra S, Bhagwat P.
Radiation-induced segregation in austenitic stainless steel type 304: Effect of high
fraction of twin boundaries. Mater Sci Eng A 2011;528:7541-51.
[325] Behjati P, Kermanpur A, Najafizadeh A, Samaei Baghbadorani H, Karjalainen LP, Jung
JG, Lee YK. Design of a new Ni-free austenitic stainless steel with unique ultrahigh
strength-high ductility synergy. Mater Des 2014;63:500-7.
[326] Shen YF, Li XX, Sun X, Wang YD, Zuo L. Twinning and martensite in a 304 austenitic
stainless steel. Mater Sci Eng A 2012;552:514-22.
[327] Tiamiyu AA, Eskandari M, Nezakat M, Wang X, Szpunar JA, Odeshi AG. A
comparative study of the compressive behaviour of AISI 321 austenitic stainless steel
under quasi-static and dynamic shock loading. Mater Des 2016;112:309-19.
[328] Eskandari M, Kermanpur A, Najafizadeh A. Formation of nanocrystalline structure in
301 stainless steel produced by martensite treatment. Metall Mater Trans A
2009;40:2241-9.
[329] Di Schino A, Salvatori I, Kenny JM. Effects of martensite formation and austenite
reversion on grain refining of AISI 304 stainless steel. J Mater Sci 2002;37:4561-5.
[330] Kim SI, Yoo YC. Dynamic recrystallization behavior of AISI 304 stainless steel. Mater
Sci Eng A 2001;311:108-13.
[331] Momeni A, Dehghani K, Keshmiri H, Ebrahimi GR. Hot deformation behavior and
microstructural evolution of a superaustenitic stainless steel. Mater Sci Eng A
2010;527:1605-11.
[332] Mandal S, Bhaduri AK, Sarma VS. A study on microstructural evolution and dynamic
recrystallization during isothermal deformation of a Ti-modified austenitic stainless
steel. Metall Mater Trans A 2011;42:1062-72.
[333] Dehghan-Manshadi A, Barnett MR, Hodgson PD. Hot deformation and recrystallization
of austenitic stainless steel: Part I. Dynamic recrystallization. Metall Mater Trans A
2008;39:1359-70.
[334] Belyakov A, Tsuzaki K, Miura H, Sakai T. Effect of initial microstructures on grain
refinement in a stainless steel by large strain deformation. Acta Mater 2003;51:847-61.
[335] Zhao J, Ding H, Jiang Z, Wei D, Linghu K. Effects of hydrogen on the critical
conditions for dynamic recrystallization of titanium alloy during hot deformation.
Metallu Mater Trans A 2014;45:4932-45.
[336] Liu XG, Zhang LG, Qi RS, Chen L, Jin M, Guo BF. Prediction of critical conditions for
dynamic recrystallization in 316LN austenitic steel. J Iron Steel Res Int 2016;23(3):238-
43.
[337] Mandal S, Sivaprasad PV, Dube RK. Kinetics, mechanism and modelling of
microstructural evolution during thermomechanical processing of a 15Cr-15Ni-2.2Mo-
Ti modified austenitic stainless steel. J Mater Sci 2007;42:2724-34.
[338] Mandal GK, Stanford N, Hodgson P, Beynon JH. Effect of hot working on dynamic
118
recrystallisation study of as-cast austenitic stainless steel. Mater Sci Eng A
2012;556:685-95.
[339] Dehghan-Manshadi A, Barnett MR, Hodgson PD. Recrystallization in AISI 304
austenitic stainless steel during and after hot deformation. Mater Sci Eng A
2008;485:664-72.
[340] Sun H, Sun Y, Zhang R, Wang M, Tang R, Zhou Z. Hot deformation behavior and
microstructural evolution of a modified 310 austenitic steel. Mater Des 2014;64:374-80.
[341] Han Y, Wu H, Zhang W, Zou D, Liu G, Qiao G. Constitutive equation and dynamic
recrystallization behavior of as-cast 254SMO super-austenitic stainless steel. Mater Des
2015;69:230-40.
[342] Samantaray D, Mandal S, Phaniraj C, Bhaduri AK. Flow behavior and microstructural
evolution during hot deformation of AISI Type 316 L(N) austenitic stainless steel.
Mater Sci Eng A 2011;528:8565-72.
[343] Arun Babu K, Mandal S, Athreya CN, Shakthipriya B, Sarma VS. Hot deformation
characteristics and processing map of a phosphorous modified super austenitic stainless
steel. Mater Des 2017;115:262-75.
[344] Yanushkevich Z, Lugovskaya A, Belyakov A, Kaibyshev R. Deformation
microstructures and tensile properties of an austenitic stainless steel subjected to
multiple warm rolling. Mater Sci Eng A 2016;667:279-85.
[345] Tikhonova M, Belyakov A, Kaibyshev R. Strain-induced grain evolution in an
austenitic stainless steel under warm multiple forging. Mater Sci Eng A 2013;564:413-
22.
[346] Belyakov A, Yanushkevich Z, Tikhonova M, Kaibyshev R. On regularities of grain
refinement through large strain deformation. Mater Sci Forum 2015;838-839:314-9.
[347] Ikeda S, Tone S, Takashima S, Kaji H. Effect of thermos-mechanical control process on
strengthening of a 22Mn-13Cr-5Ni austenitic stainless steel plate for cryogenic use. ISIJ
Int 1990;30:600-7.
[348] Han J, Wang ZY, Jiang LZ. Properties of a 304 austenitic stainless steel hot strip by
TMCP. J Iron Steel Res Int 2007;14(Supplement 1):282-7.
[349] Schino AD, Barteri M, Kenny JM. Grain size dependence of mechanical, corrosion and
tribological properties of high nitrogen stainless steels. J Mater Sci 2003;38:3257-62.
[350] Schino AD, Barteri M, Kenny JM. Effects of grain size on the properties of a low nickel
austenitic stainless steel. J Mater Sci 2003;38:4725-33.
[351] Penttilä S, Toivonen A, Li J. Effect of surface modification on the corrosion resistance
of austenitic stainless steel 316L in supercritical water conditions. J of Supercrit Fluids
2013;81:157-63.
[352] Abe H, Hong SM, Watanabe Y. Oxidation behavior of austenitic stainless steels as fuel
cladding candidate materials for SCWR in superheated steam. Nucl Eng Des
2014;280:652-60.
[353] Nezakat M, Akhiani H, Penttilä S, Sabet SM, Szpunar J. Effect of thermo-mechanical
processing on oxidation of austenitic stainless steel 316L in supercritical water. Corros
Sci 2015;94:197-206.
[354] Kim JH, Kim BK, Kim DI, Choi PP, Raabe D, Yi KW. The role of grain boundaries in
the initial oxidation behavior of austenitic stainless steel containing alloyed Cu at
700 °C for advanced thermal power plant applications. Corros Sci 2015;96:52-66.
[355] Prasad YVRK, Seshacharyulu T. Modelling of hot deformation for microstructural
control. Int Mater Rev 1998;43:243-58.
[356] Tan S, Wang Z, Cheng S, Liu Z, Han J, Fu W. Processing maps and hot workability of
Super304H austenitic heat-resistant stainless steel. Mater Sci Eng A 2009;517:312-5.
[357] Xi T, Yang C, Shahzad MB, Yang K. Study of the processing map and hot deformation
behavior of a Cu-bearing 317LN austenitic stainless steel. Mater Des 2015;87:303-12.
[358] Sun H, Sun Y, Zhang R, Wang M, Tang R, Zhou Z. Study on hot workability and
optimization of process parameters of a modified 310 austenitic stainless steel using
119
processing maps. Mater Des 2015;67:165-72.
[359] Cheng Y, Du H, Wei Y, Hou L, Liu B. Metadynamic recrystallization behavior and
workability characteristics of HR3C austenitic heat-resistant stainless steel with
processing map. J Mater Process Technol 2016;235:134-42.
[360] Samantaray D, Mandal S, Jayalakshmi M, Athreya CN, Bhaduri AK, Sarma VS. New
insights into the relationship between dynamic softening phenomena and efficiency of
hot working domains of a nitrogen enhanced 316L(N) stainless steel. Mater Sci Eng A
2014;598:368-75.
[361] Venugopal S, Mannan SL. Processing map for cold and hot working of stainless steel
type AISI 304. J Nucl Mater 1993;206:77-81.
[362] Venugopal S, Mannan SL, Prasad YVRK. Processing maps for hot working of
commercial grade wrought stainless steel type AISI 304. Mater Sci Eng A
1994;177:143-9.
[363] Belyakov A, Miura H, Sakai T. Dynamic recrystallization in ultra fine-grained 304
stainless steel. Scr Mater 2000;43:21-6.
[364] Sahu N, Selokar A, Prakash U. Thermo-mechanical behaviour of 23/8 austenitic
stainless steel. ISIJ Int 2014;54:970-8.
[365] Venugopal S, Sivaprasad PV. A journal with prasad’s processing maps. J Mater Eng
Perform 2003;12:674-86.
[366] Venugopal S, Mannan SL, Prasad YVRK. Influence of cast versus wrought
microstructure on the processing map for hot working of stainless steel type AISI 304.
Mater Lett 1993;17:388-92.
[367] Tomimura K, Takaki S, Tanimoto S, Tokunaga Y. Optimal chemical composition in Fe-
Cr-Ni alloys for ultra grain refining by reversion from deformation induced martensite.
ISIJ Int 1991;31:721-7.
[368] Naghizadeh M, Mirzadeh. Microstructural evolutions during annealing of plastically
deformed aisi 304 austenitic stainless steel: martensite reversion, grain refinement,
recrystallization, and grain growth. Metall Mater Trans A 2016;47:4210-6.
[369] Somani MC, Juntunen P, Karjalainen LP, Misra RDK, Kyröläinen A. Enhanced
mechanical properties through reversion in metastable austenitic stainless steels. Metall
Mater Trans A 2009;40:729-44.
[370] Forouzan F, Najafizadeh A, Kermanpur A, Hedayati A, Surkialiabad R. Production of
nano/submicron grained AISI 304L stainless steel through the martensite reversion
process. Mater Sci Eng A 2010;527:7334-9.
[371] Olson GB, Cohen M. Kinetics of strain-induced martensitic nucleation. Metall Trans A
1975;6:791-5.
[372] Misra RDK, Nayak S, Mali SA, Shah JS, Somani MC, Karjalainen LP. On the
significance of nature of strain-induced martensite on phase-reversion-induced
nanograined/ultrafine-grained austenitic stainless steel. Metall Mater Trans A
2010;41:3-12.
[373] Shirdel M, Hirzadeh H, Parsa MH. Nano/ultrafine grained austenitic stainless steel
through the formation and reversion of deformation-induced martensite: Mechanisms,
microstructures, mechanical properties, and TRIP effect. Mater Charact 2015;103:150-
61.
[374] Johannsen DL, Kyrolainen A, Ferreira PJ. Influence of annealing treatment on the
formation of nano/submicron grain size AISI 301 austenitic stainless steels. Metal
Mater Trans A 2006;37:2325-38.
[375] Karimi M, Najafizadeh A, Kermanpur A, Eskandari M. Effect of martensite to austenite
reversion on the formation of nano/submicron grained AISI 301 stainless steel. Mater
Charact 2009;60:1220-3.
[376] Mangonon JRPL, Thomas G. Structure and properties of thermal-mechanically treated
304 stainless steel. Metall Trans 1970;1:1587-94.
[377] Guy KB, Butler EP, West DRF. Reversion of bcc alpha prime martensite in Fe-Cr-Ni
120
austenitic stainless steels. Metal Sci 1983;17:167-76.
[378] Rajasekhara S, Karjalainen LP, Kyröläinen A, Ferreira PJ. Microstructure evolution in
nano/submicron grained AISI 301LN stainless steel. Mater Sci Eng A 2010;527:1986-
96.
[379] Misra RDK, Nayak S, Venkatasurya PKC, Ramuni V, Somani MC, Karjalainen LP.
Nanograined/ultrafine-grained structure and tensile deformation behavior of shear phase
reversion-induced 301 austenitic stainless steel. Metall Mater Trans A 2010;41:2162-74.
[380] Xu DM, Li GQ, Wan XL, Xiong RL, Xu G, Wu KM, Somani MC, Misra RDK.
Deformation behavior of high yield strength – High ductility ultrafine-grained 316LN
austenitic stainless steel. Mater Sci Eng A 2017;688:407-15.
[381] Kisko A, Hamada AS, Talonen J, Porter D, Karjalainen LP. Effects of reversion and
recrystallization on microstructure and mechanical properties of Nb-alloyed low-Ni
high-Mn austenitic stainless steels. Mater Sci Eng A 2016;657:359-70.
[382] Rajasekhara S, Ferreira PJ, Karjalainen LP, Kyröläinen A. Hall-Petch behavior in ultra-
fine-grained AISI 301LN stainless steel. Metall Mater Trans A 2007;38:1202-10.
[383] Challa VSA, Misra RDK, Somani MC, Wang ZD. Influence of grain structure on the
deformation mechanism in martensitic shear reversion-induced Fe-16Cr-10Ni model
austenitic alloy with low interstitial lcontent: Coarse-grained versus nano-
grained/ultrafine-grained structure. Mater Sci Eng A 2016;661:51-60.
[384] Behjaki P, Kermanpur A, Najafizadeh A, Samaei Baghbadorani H. Effect of annealing
temperature on nano/ultrafine grain of Ni-free austenitic stainless steel. Mater Sci Eng
A 2014;592:77-82.
[385] Baghbadorani HS, Kermanpur A, Najafizadeh A, Behjati P, Moallemi M, Rezaee A.
Influence of Nb-microalloying on the formation of nano/ultrafine-grained
microstructure and mechanical properties during martensite reversion process in a 201-
type austenitic stainless steel. Metall Mater Trans A 2015;46:3406-13.
[386] Sun GS, Du LX, Hu J, Xie H, Wu HY, Misra RDK. Ultrahigh strength nano/ultrafine-
grained 304 stainless steel through three-stage cold rolling and annealing treatment.
Mater Charact 2015;110:228-35.
[387] Rezaee A, Najafizadeh A, Kermanpur A, Moallemi M. The influence of reversion
annealing behavior on the formation of nanograined structure in AISI 201L austenitic
stainless steel through martensite treatment. Mater Des 2011;32:4437-42.
[388] Hall JN, Fekete JR. Steels for auto bodies: A general overview. In: Rana R, Singh SB
(eds.). Automotive Steels: Design, Metallurgy, Processing and Applications. Elsevier;
2017. pp. 19-45.
[389] Rana R. Low-density steels. JOM 2014;66:1730-33.
[390] Vercammen S, Blanpain B, De Cooman BC, Wollants P. Cold rolling behaviour of an
austenitic Fe-30Mn-3Al-3Si TWIP-steel: the importance of deformation twinning. Acta
Mater 2004;52:2005-12.
[391] Benito JA, Cobo R, Lei W, Calvo J, Cabrera JM. Stress-strain response and
microstructural evolution of a FeMnCAl TWIP steel during tension-compression tests.
Mater Sci Eng A 2016;655:310-20.
[392] Sohn SS, Song H, Suh BC, Kwak JH, Lee BJ, Kim NJ, Lee S. Novel ultra-high-strength
(ferrite + austenite) duplex lightweight steels achieved by fine dislocation substructures
(Taylor lattices), grain refinement, and partial recrystallization. Acta Mater
2015;96:301-10.
[393] Yao MJ, Dey P, Seol JB, Choi P, Herbig M, Marceau RKW, Hickel T, Neugebauer J,
Raabe D. Combined atom probe tomography and density functional theory investigation
of the Al off-stoichiometry of κ-carbides in an austenitic Fe-Mn-Al-C low density steel.
Acta Mater 2016;106:229-38.
[394] Yoo JD, Park KT. Microband-induced plasticity in a high Mn-Al-C light steel. Mater
Sci Eng A 2008;496:417-24.
[395] Song W, zhang W, Appen J, Dronskowski R, Bleck W. κ-phase formation in Fe-Mn-Al-
121
C austenitic steels. Steel Res Int 2015;86:1161-9.
[396] Frommeyer G, Brüx U. Microstructures and mechanical properties of high-strength Fe-
Mn-Al-C light-weight TRIPLEX steels. Steel Res Int 2006;77:627-33.
[397] Rana R, Liu C, Ray RK. Low-density low-carbon Fe-Al ferritic steels. Scr Mater
2013;68:354-9.
[398] Sohn SS, Lee BJ, Lee S, Kim NJ, Kwak JH. Effect of annealing temperature on
microstructural modification and tensile properties in 0.35 C- 3.5Mn-5.8 Al lightweight
steel. Acta Mater 2013;61:5050-66.
[399] Wang J, Wang Z, Wang X, Yang Q, Jin X, Wang L. Strengthening effect of nanoscale
precipitation and transformation induced plasticity in a hot rolled copper-containing
ferrite-based lightweight steel. Scr Mater 2017:129:25-9.
[400] Kubaschewski O. Iron-Binary Phase Diagrams. Springer; 1982.
[401] Pierce DT, Jiménez JA, Bentley J, Raabe D, Wittig JE. The influence of stacking fault
energy on the microstructural and strain-hardening evolution of Fe–Mn–Al–Si steels
during tensile deformation. Acta Mater 2015;100:178-90.
[402] Han SY, Shin SY, Lee S, Kim NJ, Kwak JH, Chin KG. Effect of carbon content on
cracking phenomenon occurring during cold rolling of three light-weight steel plates.
Metall Mater Trans A 2011;42:138-46.
[403] Seo CH, Kwon KH, Choi K, Kim KH, Kwak JH, Lee S, Kim NJ. Deformation behavior
of ferrite-austenite duplex lightweight Fe-Mn-Al-C steel. Scr Mater 2012;66:519-22.
[404] Kim SH, Kim H, Kim NJ. Brittle intermetallic compound makes ultrastrong low-density
steel with large ductility. Nature 2015;518:77-9.
[405] Yang HK, Zhang ZJ, Zhang ZF. Comparison of work hardening and deformation
twinning evolution in Fe-22Mn-0.6C-(1.5Al) twinning-induced plasticity steels. Scr
Mater 2013;68:992-5.
[406] Boyer HE, Gall TL. Metals Handbook: Desk Edition. American Society for Metals,
Metals Park, Ohio; 1985.
[407] Hamada AS. Manufacturing, Mechanical Properties and Corrosion Behaviour of High-
Mn TWIP Steels. Oulu University Press; 2007.
[408] Chen S, Rana R, Haldar A, Ray RK. Current states of Fe-Mn-Al-C low density steels.
Prog Mater Sci 2017;89:345-91.
[409] Kim H, Suh DW, Kim NJ. Fe–Al–Mn–C lightweight structural alloys: a review on the
microstructures and mechanical properties. Sci Tech Adv Mater 2013;14:1–11.
[410] Zuazo I, Hallstedt B, Lindahl B, Selleby M, Soler M, Etienne A, Perlade A, Hasenpouth
D, Massardier-Jourdan V, Cazottes S, Kleber X. Low-density steels: complex
metallurgy for automotive applications. JOM 2014;66:1747-58.
[411] Rana R, Lahaye C, Ray RK. Overview of lightweight ferrous materials: Strategies and
promises. JOM 2014;66:1734-46.
[412] Lu WJ, Qin RS. Influence of κ-carbide interface structure on the formability of
lightweight steels. Mater Des 2016;104:211-6.
[413] Rana R, Liu C, Ray RK. Evolution of microstructure and mechanical properties during
thermomechanical processing of a low-density multiphase steel for automotive
application. Acta Mater 2014;75:227-45.
[414] Zargaran A, Kim HS, Kwak JH, Kim NJ. Effect of C content on the microstructure and
tensile properties of lightweight ferritic Fe-8Al-5Mn-0.1Nb alloy. Met Mater Int
2015;21:79-84.
[415] Wang J, Yang Q, Wang X, Wang L. Effect of coiling temperature on microstructure and
tensile behavior of a hot-rolled ferritic lightweight steel. Metall Mater Trans A
2016;47:5918-31.
[416] Shin SY, Lee H, Han SY, Seo CH, Choi K, Lee S, Kim NJ, Kwak JH, Chin KG.
Correlation of microstructure and cracking phenomenon occurring during hot rolling of
lightweight steel plates. Metall Mater Trans A 2010;41:138-48.
[417] Abedi HR, Zarei Hanzaki A, Liu Z, Xin R, Haghdadi N, Hodgson PD. Continuous
122
dynamic recrystallization in low density steel. Mater Des 2017;114:55-64.
[418] Mohamadizadeh A, Zarei-Hanzaki A, Kisko A, Porter D. Ultra-fine grained structure
formation through deformation-induced ferrite formation in duplex lowdensity steel.
Mater Des 2016;92:322-9.
[419] Mohamadizadeh A, Zarei-Hanzaki A, Abedi HR. Modified constitutive analysis and
activation energy evolution of a low-density steel considering the effects of deformation
parameters. Mech Mater 2016;95:60-70.
[420] Sohn SS, Lee BJ, Lee S, Kwak JH. Effects of aluminum content on cracking
phenomenon cracking during cold rolling of three ferrite-based lightweight steel. Acta
Mater 2013;61:5626-35.
[421] Rana R, Loiseaux J, Lahaye C. Microstructure, mechanical properties and formability of
a duplex steel. Mater Sci Forum 2012;706-709:2271-7.
[422] Sohn SS, Choi K, Kwak JH, Kim NJ, Lee S. Novel ferrite-austenite duplex lightweight
steel with 77% ductility by transformation induced plasticity and twinning induced
plasticity mechanisms. Acta Mater 2014;78:181-9.
[423] Sohn SS, Song H, Kim JG, Kwak JH, Kim HS, Lee S. Effects of annealing treatment
prior to cold rolling on delayed fracture properties in ferrite-austenite duplex
lightweight steels. Metall Mater Trans A 2016;47:706-17.
[424] Song H, Lee SG, Sohn SS, Kwak JH, Lee S. Effect of strain-induced age hardening on
yield strength improvement in ferrite-austenite duplex lightweight steels. Metall Mater
Trans A 2016;47:5372-82.
[425] Yang F, Song R, Li Y, Sun T, Wang K. Tensile deformation of low density duplex Fe-
Mn-Al-C steel. Mater Des 2015;76:32-9.
[426] Rana R, Liu C. Effects of ceramic particles and composition on elastic modulus of low
density steels for automotive applications. Can Metall Quarterly 2014;53:300-16.
[427] Hwang SW, Ji JH, Lee EG, Park KT. Tensile deformation of a duplex Fe-20Mn-9Al-
0.6C steel having the reduced specific weight. Mater Sci Eng A 2011;528:5196-203.
[428] Sohn SS, Lee BJ, Kwak JH, Lee S. Effects of annealing treatment prior to cold rolling
on the edge cracking phenomenon of ferritic lightweight steel. Metall Mater Trans A
2014;45:3844-56.
[429] Zargaran A, Kim HS, Kwak JH, Kim NJ. Effects of Nb and C additions on the
microstructure and tensile properties of lightweight ferritic Fe-8Al-5Mn alloy. Scr
Mater 2014;89:37-40.
[430] Lee CY, Jeong J, Han J, Lee SJ, Lee S, Lee YK. Coupled strengthening in a medium
manganese lightweight steel with an inhomogeneously grained structure of austenite.
Acta Mater 2015;84:1-8.
[431] Chio K, Seo CH, Lee H, Kim SK, Kwak JH, Chin KG, Park KT, Kim NJ. Effect of
aging on the microstructure and deformation behavior of austenite base lightweight Fe
–28Mn–9Al–0.8C steel. Scr Mater 2010;63:1028-31.
[432] Song H, Sohn SS, Kwak JH, Lee BJ, Lee S. Effect of austenite stability on
microstructural evolution and tensile properties in intercritically annealed medium-Mn
lightweight steels. Metall Mater Trans A 2016;47:2674-85.
[433] Han SY, Shin SY, Lee HJ, Lee BJ, Lee S, Kim NJ, Kwak JH. Effects of annealing
temperature on microstructure and tensile properties in ferritic lightweight steels. Metall
Mater Trans A 2012;43:843-53.
[434] Abedi HR, Hanzaki AZ, Ou KL, Yu CH. Substructure hardening in duplex low density
steel. Mater Des 2017;116:472-80.
[435] Ha MC, Koo JM, Lee JK, Hwang SW, Park KT. Tensile deformation of a low density
Fe-27Mn-12Al-0.8C duplex steel in associate with ordered phases at ambient
temperature. Mater Sci Eng A 2013;586:276-83.
[436] Lu WJ, Zhang XF, Qin RS. κ-carbide hardening in a low-density high-Al high-Mn
multiphase steel. Mater Lett 2015;138:96-9.

123
[437] Park SJ, Hwang B, Lee KH, Lee TH, Suh DW, Han HN. Microstructure and tensile
behavior of duplex low-density steel containing 5 mass% aluminum. Scr Mater
2013;68:365-9.
[438] Haase C, Zehnder C, Ingendahl T, Bikar A, Tang F, Hallstedt B, Hu W, Bleck W,
Molodov DA. On the deformation behavior of k-carbide-free and k-carbide-containing
high-Mn ligh-weight steel. Acta Mater 2017;122:332-43.
[439] Lee K, Park SJ, Lee J, Moon J, Kang JY, Kim DI, Suh JY, Han HN. Effect of aging
treatment on microstructure and intrinsic mechanical behavior of Fe-31.4Mn-11.4Al-
0.89C lightweight steel. J Alloy Compd 2016;656:805-11.
[440] Kim SK, Kim G, Chin KG. Development of high manganese TWIP steel with 980MPa
tensile strength. Proceeding of the International Conference on New Developments in
Advanced High-Strength Sheet Steels. AIST, Orlando, USA; 15-18 June 2008. pp. 249-
56.
[441] Matlock DK, Speer JG. Third Generation of AHSS: Microstructure Design Concepts. In:
Arunansu H (ed.). Microstructure and Texture in Steels and Other Materials. Springer;
2009. pp. 185-205.
[442] Matlock DK, Speer JG, De Moor E, Gibbs PJ. Resent developments in advanced high
strength sheet steels for automotive applications: an overview. JESTECH 2012;15(1):1-
12.
[443] Olson GB, Cohen M. Kinetics of strain-induced martensite nucleation. Metall Trans A
1975;6A:791-5.
[444] Qu H. Advanced High Strength Steel Through Paraequilibrium Carbon Partitioning and
Austenite Stabilization (PhD Thesis). Case Western Reserve University, Cleveland;
2013.
[445] Qu H. Advanced High Strength Steel Through Paraequilibrium Carbon Partitioning and
Austenite Stabilization (Master Thesis). Case Western Reserve University, Cleveland;
2011.
[446] Gibbs PJ. Design Considerations for the Third Generation Advanced High Strength
Steel. Colorado School of Mines, USA; 2012.
[447] Poliak EI, Bhattacharya D. Aspects of thermomechanical processing of 3 rd generation
advanced high strength steels. Mater Sci Forum 2014;783-786:3-8.
[448] Askari-Paykani M, Shahverdi HR, Miresmaeili R. First and third generations of
advanced high-strength steels in a FeCrNiBSi system. J Mater Process Technol
2016;238:383-94.
[449] De Moor E, Gibbs PJ, Speer JG, Matlock DK, Schroth JG. Strategies for third-
generation advanced high-strength steel development. AIST Trans 2010;7(3):133-44.
[450] Speer JG, Matlock DK, De Cooman BC, Schroth JG. Carbon partitioning into austenite
after martensite transformation. Acta Mater 2003;51:2611-22.
[451] Toji Y, Miyamoto G, Raabe D. Carbon partitioning during quenching and partitioning
heat treatment accompanied by carbide precipitation. Acta Mater 2015;86:137-47.
[452] Zhang J, Ding H, Wang C, Zhao J, Ding T. Work hardening behaviors of a low carbon
Nb-microalloyed Si-Mn quenching-partitioning steel with different cooling styles after
partitioning. Mater Sci Eng A 2013;585:132-8.
[453] Santofimia MJ, Zhao L, Petrov R, Kwakernaak C, Sloof WG, Sietsma J.
Microstructural development during the quenching and partitioning process in a newly
designed low-carbon steel. Acta Mater 2011;59:6059-68.
[454] Speer JG, De Moor E, Findley KO, Matlock DK, De Cooman BC, Edmonds DV.
Analysis of microstructure evolution in quenching and partitioning automotive sheet
steel. Metall Mater Trans A 2011;42:3591-601.
[455] Thomas GA, Speer JG, Matlock DK. Quenched and partitioned microstructures
produced via Gleeble simulations of hot-strip mill cooling practices. Metall Mater Trans
A 2011;42:3652-9.
[456] Liu H, Lu X, Jin X, Dong H, Shi J. Enhanced mechanical properties of a hot stamped
124
advanced high-strength steel treated by quenching and partitioning process. Scr Mater
2011;64:749-52.
[457] Speer J, Matlock DK, De Cooman BC, Schroth JG. Carbon partitioning into austenite
after martensite transformation. Acta Mater 2003;51:2611-22.
[458] Toji Y, Matsuda H, Herbig M, Choi PP, Raabe D. Atomic-scale analysis of carbon
partitioning between martensite and austenite by atom probe tomography and
correlative transmission electron microscopy. Acta Mater 2014;65:215-28.
[459] Zhang K, Liu P, Li W, Guo Z, Rong Y. Ultrahigh strength-ductility steel treated by a
novel quenching-partitioning-tempering process. Mater Sci Eng A 2014;619:205-11.
[460] Arlazarov A, Bouazia O, Masse JP, Kegel F. Characterization and modeling of
mechanical behavior of quenching and partitioning steels. Mater Sci Eng A
2015;620:293-300.
[461] Bhadeshia HKDH, Edmonds DV. The bainite transformation in a silicon steel. Metall
Trans A 1979;10:895-907.
[462] Koistinen DP, Marburger RE. A general equation prescribing the extent of the
austenite-martensite transformation in pure iron-carbon alloys and plain carbon steels.
Acta Metall 1959;7:59-60.
[463] Wang XD, Zhong N, Rong YH, Hsu TY, Wang L. Novel ultrahigh-strength nanolath
martensitic steel by quenching-partitioning-tempering process. J Mater Res
2009;24:260-7.
[464] Edmonds DV, He K, Rizzo FC, De Cooman BC, Matlock DK, Speer JG. Quenching
and partitioning martensite-A novel steel heat treatment. Mater Sci Eng A 2006;438-
440:25-34.
[465] De Knijf D, Petrov R, Föjer C, Kestens LAI. Effect of fresh martensite on the stability
of retained austenite in quenching and partitioning steels. Mater Sci Eng A
2014;615:107-15.
[466] Santofimia MJ, Nguyen-Minh T, Zhao L, Petrov R, Sabirov I, Sietsma J. New low
carbon Q&P steels containing film-like intercritical ferrite. Mater Sci Eng A
2010;527:6429-39.
[467] Zhang J, Ding H, Misra RDK. Enhanced strain hardening and microstructural
characterization in a low carbon quenching and partitioning steel with partial
austenization. Mater Sci Eng A 2015;636:53-9.
[468] Speer JG, Rizzo FC, Matlock DK, Edmonds DV. The “quenching and partitioning”
process: background and recent progress. Mater Res 2005;8:417-23.
[469] Santofimia MJ, Zhao L, Sietsma J. Overview of mechanisms involved during the
quenching and partitioning process in steels. Metall Mater Trans A 2011;42:3620-6.
[470] Speer JG, Streicher AM, Matlock DK, Rizzo FC, Krauss G. Quenching and partitioning:
a fundamentally new process to create high strength TRIP sheet microstructures. In:
Damm EB, Merwin M (eds.). Austenite Formation and Decomposition. Warrendale, PA:
TMS/ISS; 2003. pp. 505-522.
[471] Tan X, Xu Y, Yang X, Liu Z, Wu D. Effect of partitioning procedure on microstructure
and mechanical properties of a hot-rolled directly quenched and partitioned steel. Mater
Sci Eng A 2014;594:149-60.
[472] Matsumura O, Sakuma Y, Takechi H. Enhancement of elongation by retained austenite
in intercritical annealed 0.4C-1.5Si-O.8Mn steel. Trans Iron Steel Inst Jpn 1987;27:570-
9.
[473] Sakuma Y, Matsumura O, Takechi H. Enhancement of elongation by retained austenite
in intercritical annealed 0.4C-1.5Si-O.8Mn steel. Metall Trans A 1991;22:489-98.
[474] Lee SJ, Lee S, De Cooman BC. Mn partitioning during the intercritical annealing of
ultrafine-grained 6% Mn transformation-induced plasticity steel. Scr Mater
2011;64:649-52.
[475] Izumiyama M, Tsuchiya M, Imai Y.

125
J Jpn Inst Met
1970;22:105-15.
[476] Jimenez-Melero E, Van Dijk NH, Zhao L, Sietsma J, Offerman SE, Wright JP, van der
Zwaag S. Martensitic transformation of individual grains in low-alloyed TRIP steels.
Scr Mater 2007;56:421-4.
[477] Breedis JF. Influence of dislocation substructure on the martensitic transformation in
stainless steel. Acta Metall 1965;13:239-50.
[478] Luo H, Shi J, Wang C, Cao W, Sun X, Dong H. Experimental and numerical analysis
on formation of stable austenite during the intercritical annealing of 5Mn steel. Acta
Mater 2011;59:4002-14.
[479] Shi J, Sun X, Wang M, Hui W, Dong H, Cao W. Enhanced work-hardening behavior
and mechanical properties in ultrafine-grained steels with large-fractioned metastable
austenite. Scr Mater 2010;63:815-8.
[480] Nakada N, Arakawa Y, Park KS, Tsuchiyama T, Takaki S. Dual phase structure formed
by partial reversion of cold-deformed martensite. Mater Sci Eng A 2012;553:128-33.
[481] Emadoddin E, Akbarzadeh A, Daneshi G. Effect of intercritical annealing on retained
austenite characterization in textured TRIP-assisted steel sheet. Mater Charact
2006;57:408-13.
[482] Hu B, Luo H, Yang F, Dong H. Recent progress in medium-Mn steels made with new
designing strategies, a review. J Mater Sci Techol 2017. DOI:
http://dx.doi.org/10.1016/j.jmst.2017.06.017.
[483] Zhou WH, Wang XL, Venkatsurya PKC, Guo H, Shang CJ, Misra RDK. Structure-
mechanical property relationship in a highstrength low carbon alloy steel processed by
two-step intercritical annealing and intercritical tempering. Mater Sci Eng A
2014;607:569-77.
[484] Raabe D, Sandlöbes S, Millán J, Ponge D, Assadi H, Herbig M, Choi PP. Segregation
engineering enables nanoscale martensite to austenite phase transformation at grain
boundaries: A pathway to ductile martensite. Acta Mater 2013;61:6123-52.
[485] Kuzmina M, Ponge D, Raabe D. Grain boundary segregation engineering and austenite
reversion turn embrittlement into toughness: Example of a 9 wt.% medium Mn steel.
Acta Mater 2015;86:182-92.
[486] Wang MM, Tasan CC, Ponge D, Dippel ACh, Raabe D. Nanolaminate transformation-
induced plasticity–twinning-induced plasticity steel with dynamic strain partitioning
and enhanced damage resistance. Acta Mater 2015;85:216-28.
[487] Tsuchiyama T, Inoue T, Tobata J, Akama D, Takaki S. Microstructure and mechanical
properties of a medium manganese steel treated with interrupted quenching and
intercritical annealing. Scr Mater 2016;122:36-9.
[488] Lee SW, Lee KY, De Cooman BC. Ultra fine-grained 6wt% manganese TRIP steel.
Mater Sci Forum 2010;654-656:286-9.
[489] Springer H, Belde M, Raabe D. Bulk combinatorial design of ductile martensitic
stainless steels through confined martensite-to-austenite reversion. Mater Sci Eng A
2013;582;235-44.
[490] Raabe D, Ponge D, Dmitrieva O, Sander B. Nanoprecipitate-hardened 1.5 GPa steel
with unexpected high ductility. Scr Mater 2009;60:1141-4.
[491] Hu J, Du LX, Sun GS, Xie H, Yi HL, Misra RDK. The determining role of reversed
austenite in enhancing toughness of a novel ultra-low carbon medium manganese high
strength steel. Scr Mater 2015;104:87-90.
[492] Chen J, Lv M, Tang S, Liu Z, Wang G. Correlation between mechanical properties and
retained austenite characteristics in a low-carbon medium manganese alloyed steel plate.
Mater Charact 2015;106:108-11.
[493] Luo H, Dong H, Huang M. Effect of intercritical annealing on the Lüders strains of
medium Mn transformation-induced plasticity steels. Mater Des 2015;83:42-8.

126
[494] Zeytin HK, Kubilay C, Aydin H. Investigation of dual phase transformation of
commercial low alloy steels: Effect of holding time at low inter-critical annealing
temperatures. Mater Lett 2008;62:2651-3.
[495] Zhang Z, Li Y, Kong X, Manabe K, Wang N. Effect of intercritical annealing time on
microstructure and axial mechanical properties of TRIP seamless steel tube. Steel Res
Int 2014;85:632-9.
[496] Arlazarov A, Gouné M, Bouaziz O, Hazotte A, Kegel F. Effect of intercritical annealing
time on microstructure and mechanical behavior of advanced medium Mn steels. Mater
Sci Forum 2012;706-709:2693-8.
[497] Huang H, Matsumura O, Furukawa T. Retained austenite in low carbon, manganese
steel after intercritical heat treatment. Mater Sci Technol 1994;10:621-6.
[498] Li ZC, Ding H, Misra RDK, Cai ZH. Microstructure-mechanical property relationship
and austenite stability in medium-Mn TRIP steels: The effect of austenite-reverted
transformation and quenching-tempering treatments. Mater Sci Eng A 2017;682:211-9.
[499] Dmitrieva O, Choi P, Gerstl SSA, Ponge D, Raabe D. Pulsed-laser atom probe studies
of a precipitation hardened maraging TRIP steel. Ultramicroscopy 2011;111:623-7.
[500] Servant C, Bouzid N. Influence of the increasing content of Mo on the precipitation
phenomena occurring during tempering in the maraging alloy Fe-12Mn-9Co-5Mo. Acta
Metall 1988;36:2771-8.
[501] Tewari R, Mazumder S, Batra IS, Dey GK, Banerjee S. Precipitation in 18 wt%
maraging steel grade 350. Acta Mater 2000;48:1187-200.
[502] Kim JI, Kim HJ, Morris JW. The role of the constituent phases in determining the low
temperature toughness of 5.5Ni cryogenic steel. Metall Trans A 1984;15:2213-9.
[503] Fultz B, Morris JW. The mechanical stability of precipitated austenite in 9Ni steel.
Metall Trans A 1985;16:2251-6.
[504] Hu J, Du LX, Liu H, Sun GS, Xie H, Yi HL, Misra RDK. Structure–mechanical
property relationship in a low-C medium-Mn ultrahigh strength heavy plate steel with
austenite-martensite submicro-laminate structure. Mater Sci Eng A 2015;647:144-51.
[505] Zhang J, Ding H, Misra RDK, Wang C. Enhanced stability of retained austenite and
consequent work hardening rate through pre-quenching prior to quenching and
partitioning in a Q-P microalloyed steel. Mater Sci Eng A 2014;611:252-6.
[506] Luo H, Dong H. New ultrahigh-strength Mn-alloyed TRIP steels with improved
formability manufactured by intercritical annealing. Mater Sci Eng A 2015;626:207-12.
[507] Zhu J, Ding R, He J, Yang Z, Zhang C, Chen H. A cyclic austenite reversion treatment
for stabilizing austenite in the medium manganese steels. Scr Mater 2017;136:6-10.
[508] Wang MM, Tasan CC, Ponge D, Raabe D. Spectral TRIP enables ductile 1.1 GPa
martensite. Acta Mater 2016;111:262-72.
[509] Inoue J, Nambu S, Ishimoto Y, Koseki T. Fracture elongation of brittle/ductile
multilayered steel composites with strong interface. Scr Mater 2008;59:1055-8.
[510] Jeong C, Oya T, Yanagimoto J. Analysis of fracture behavior and stress-strain
distribution of martensite/austenite multilayered metallic sheet. J Mater Process Technol
2013;213:614-20.
[511] Wang MM, Tasan CC, Ponge D, Kostka A, Raabe D. Smaller is less stable: Size effects
on twinning vs. transformation of reverted austenite in TRIP-maraging steels. Acta
Mater 2014;79:268-81.
[512] Koyama M, Zhang Z, Wang M, Ponge D, Raabe D, Tsuzaki K, Noguchi H, Tasan CC.
Bone-like crack resistance in hierarchical metastable nanolaminate steels. Sci
2017;355:1055-7.
[513] Hsu TY, Xu ZY. Design of structure, composition and heat treatment process for high
strength steel. Mater Sci Forum 2007;561-565:2283-6.
[514] Hsu TY, Jin XJ, Rong YH. Strengthening and toughening mechanisms of quenching-
partitioning-tempering (Q-P-T) steels. J Alloy Compd 2013;577S:S568-71.
[515] Wang Y, Guo Z, Chen N, Rong Y. Deformation temperature dependence of mechanical
127
properties and microstructures for a novel quenching-partitioning-tempering steel. J
Mater Sci Technol 2013;29:451-7.
[516] Gao G, Zhang H, Gui X, Tan Z, Bai B. Tempering behavior of ductile 1700 MPa Mn–
Si–Cr–C Steel treated by quenching and partitioning process incorporating bainite
formation. J Mater Sci Technol 2015;31:199-204.
[517] Wang XD, Guo WZ, Wang L, Rong YH. Carbide characterization in a Nb-microalloyed
advanced ultrahigh strength steel after quenching-partitioning-tempering process. Mater
Sci Eng A 2010;527:3373-8.
[518] Zhang K, Zhang M, Guo Z, Chen N, Rong Y. A new effect of retained austenite on
ductility enhancement in high-strength quenching–partitioning–tempering martensitic
steel. Mater Sci Eng A 2011;528:8486-91.
[519] Wang XD, Guo ZH, Rong YH. Mechanism exploration of an ultrahigh strength steel by
quenching-partitioning-tempering process. Mater Sci Eng A 2011;529:35-40.
[520] Qu H, Michal GM, Heuer AH. A 3rd generation advanced high-strength steel (AHSS)
produced by dual stabilization heat treatment (DSHT). Metall Mater Trans A
2013;44:4450-3.
[521] Qu H, Michal GM, Heuer AH. Third generation 0.3C-4.0Mn advanced high strength
steels through a dual stabilization heat treatment: austenite stabilization through
paraequilibrium carbon partitioning. Metall Mater Trans A 2014;45:2741-9.
[522] Huang X, Liu W, Huang Y, Chen H, Huang W. Effect of a quenching-long partitioning
treatment on the microstructure and mechanical properties of a 0.2C% bainitic steel. J
Mater Process Technol 2015;222:181-7.
[523] Seto K, Matsuda H. Application of nanoengineering to research and development and
production of high strength steel sheets. Mater Sci Technol 2013;29:1158-65.
[524] Tsuyama S, Endo S, Seto K. Social contribution through JFE steel’s environment-
friendly steel products. JFE Tech Rep 2014;19:91-102.
[525] Funakawa Y, Shiozaki T, Tomita K, Yamamoto T, Maeda E. Development of high
strength hot-rolled sheet steel consisting of ferrite and nanometer-sized carbides. ISIJ
Int 2004;44:1945-51.
[526] Krauss G. Steels: Processing, Structure, and Performance. ASM International, 2005.
[527] Mukherjee S, Timokhina I, Zhu C, Ringer SP, Hodgson PD. Clustering and
precipitation processes in a ferritic titanium-molybdenum microalloyed steel. J Alloys
Compd 2017;690:621-32.
[528] Yen HW, Huang CY, Yang JR. Characterization of interphase-precipitated nanometer-
sized carbides in a Ti-Mo-bearing steel. Scr Mater 2009;61:616-9.
[529] Timokhina IB, Hodgson PD, Ringer SP, Zheng RK, Pereloma EV. Precipitate
characterisation of an advanced high-strength low-alloy (HSLA) steel using atom probe
tomography. Scr Mater 2007;56:601-4.
[530] Chen CY, Yen HW, Kao FH, Li WC, Huang CY, Yang JR, Wang SH. Precipitation
hardening of high-strength low-alloy steels by nanometer-sized carbides. Mater Sci Eng
A 2009;499;162-166.
[531] Seto K, Funakawa Y, Kaneko S. Hot rolled high strength steels for suspension and
chassis parts “NANOHITEN” and “BHT® steel”. JFE Tech Rep 2007;10:19-25.
[532] Chen CY, Yang JR, Chen CC, Chen SF. Microstructural characterization and
strengthening behavior of nanometer sized carbides in Ti–Mo microalloyed steels
during continuous cooling process. Mater Charact 2016;114:18-29.
[533] Jha G, Das S, Sinha S, Lodh A, Haldar A. Design and development of precipitate
strengthened advanced high strength steel for automotive application. Mater Sci Eng A
2013;561:394-402.
[534] Yen HW, Chen PY, Huang CY, Yang JR. Interphase precipitation of nanometer-sized
carbides in a titanium-molybdenum-bearing low-carbon steel. Acta Mater
2011;59:6264-74.

128
[535] Honeycombe RWK, Mehl RF. Transformation from austenite in alloy steels. Metall
Trans A 1976;7:915-36.
[536] Park JS, Ha YS, Lee SJ, Lee YK. Dissolution and precipitation kinetics of Nb(C,N) in
austenite of a low-carbon Nb-microalloyed steel. Metall Mater Trans A 2009;40:560-8.
[537] Kamibayashi K, Tanabe Y, Takemoto Y, Shimizu I, Senuma T. Influence of Ti and Nb
on the strength–ductility–hole expansion ratio balance of hot-rolled low-carbon high-
strength steel sheets. ISIJ Int 2012;52:151-7.
[538] Cheng L, Cai QW, Xie BS, Ning Z, Zhou XC, Li GS. Relationships among
microstructure, precipitation and mechanical properties in different depths of Ti–Mo
low carbon low alloy steel plate. Mater Sci Eng A 2016;651:185-191.
[539] Phaniraj MP, Shin YM, Lee J, Goo NH, Kim DI, Suh JY, Jung WS, Shim JH, Choi IS.
Development of high strength hot rolled low carbon copper-bearing steel containing
nanometer sized carbides. Mater Sci Eng A 2015;633:1-8.
[540] Huang Y, Zhao A, Wang X, Wang X, Yang J, Han J, Yang F. A high-strength high-
ductility Ti- and Mo-bearing ferritic steel. Metall Mater Trans A 2016;47:450-60.
[541] Park DB, Huh MY, Shim JH, Suh JY, Lee KH, Jung WS. Strengthening mechanism of
hot rolled Ti and Nb microalloyed HSLA steels containing Mo and W with various
coiling temperature. Mater Sci Eng A 2013;560:528-34.
[542] Ning Z, Cai Q, Xie B, Dong E. The effect of deformation-induced-ferrite-
transformation on nanometer-sized carbides in Ti-Mo ferritic steel. Mater Sci Technol
2017;33:1215-23.
[543] Zhang K, Li Z, Wang Z, Sun X, Yong Q. Precipitation behavior and mechanical
properties of hot-rolled high strength Ti–Mo-bearing ferritic sheet steel: The great
potential of nanometer-sized (Ti, Mo)C carbide. J Mater Res 2016;31:1254-63.
[544] Dutta B, Palmiere EJ, Sellars CM. Modelling the kinetics of strain induced precipitation
in Nb microalloyed steels. Acta Mater 2001;49:785-94.
[545] Kim YW, Kim JH, Hong SG, Lee CS. Effects of rolling temperature on the
microstructure and mechanical properties of Ti-Mo microalloyed hot-rolled high
strength steel. Mater Sci Eng A 2014;605:244-52.
[546] Kim YW, Hong SG, Huh YH, Lee CS. Role of rolling temperature in the precipitation
hardening characteristics of Ti-Mo microalloyed hot-rolled high strength steel. Mater
Sci Eng A 2014;615:255-61.
[547] Dong H, Sun X. Deformation induced ferrite transformation in low carbon steels. Curr
Opin Solid State Mater Sci 2005;9:269-76.
[548] Hall EO. The deformation and ageing of mild steel: III discussion of results. Proc Phys
Soc B 1951;64:747-53.
[549] Petch NJ. The cleavage strength of polycrystals. J Iron Steel Inst 1953;174:25-8.
[550] Ashby MF. Results and consequences of a recalculation of the frank-read and the
orowan stress. Acta Metall 1966;14:679-81.
[551] Orowan E. Classification and nomenclature of internal stresses. In: Proceedings of
Symposium on Internal Stresses. Institute of Metals; 1948. pp. 47-59.
[552] Wang X, Zhao A, Zhao Z, Huang Y, Geng Z, Yu Y. Precipitation strengthening by
nanometer-sized carbides in hot-rolled ferritic steels. J Iron Steel Res Int 2014;21:1140-
6.

129

S-ar putea să vă placă și