You are on page 1of 10

Materials Science & Engineering A 669 (2016) 437–446

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

On the work-hardening behaviour of a high manganese TWIP steel


at different deformation temperatures
Vadim Shterner, Ilana B. Timokhina, Hossein Beladi n
Institute for Frontier Materials, Deakin University, Geelong, Victoria 3216, Australia

art ic l e i nf o a b s t r a c t

Article history: In the current study, the work-hardening behaviour of a high manganese TWIP steel was investigated at
Received 17 December 2015 different deformation temperatures. At room temperature, the steel exhibited an excellent combination
Received in revised form of mechanical properties due to a unique work-hardening behaviour. There were four distinct stages
20 April 2016
observed in the work-hardening behaviour as a result of complex dynamic strain induced micro-
Accepted 25 May 2016
Available online 27 May 2016
structural reactions consisted of dynamic recovery, dislocation dissociation, stacking fault formation,
mechanical twining and dynamic strain aging. An increase in the deformation temperature significantly
Keywords: influenced the microstructure evolution, resulting in a remarkable alteration in the work-hardening
Work-hardening behaviour behaviour. Consequently, the mechanical properties of the TWIP steel were gradually deteriorated with
Temperature
the deformation temperature. The mechanical twins appeared to have a restricted influence on the work-
Mechanical twinning
hardening behaviour of the TWIP steel at room temperature and remarkably diminished with the
TWIP steel
Stacking faults temperature. The enhanced work-hardening behaviour was mostly attributed to the interaction of glide
Dynamic strain aging dislocations with stacking faults.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction experiment revealed that mechanical twinning contribution to


flow stress is very limited and the enhanced work-hardening can
Development of new materials to improve the strength-ducti- be attributed to the forest hardening resulted by extensive dis-
lity balance led to a recent development of high manganese location evolution.
TWinning Induced Plasticity (TWIP) steels [1–3]. This type of steel In a number of publications [4,35], the DSA effect was con-
can offer  1 GPa ultimate tensile strength combined with  50% sidered as a main contributor to the enhanced work-hardening
elongation at room temperature [1–3]. These exceptional me- behaviour of the TWIP steel. The DSA effect occurs only in the
chanical properties at room temperature are due to the remark- explicit temperature range, and the work-hardening behaviour
able work-hardening behaviour of the TWIP steel [1–5]. decreases at the temperatures above and below that range [4,35].
The work-hardening of TWIP steel is mostly attributed to the It has also been shown [35] that a short range reorientation of the
nano-scale mechanical twins formed in austenitic microstructure C-Mn couples, which tend to pin the dislocations and prevent
during straining [5–10]. Mechanical twins contribute to the plastic them from gliding, can contribute to the work-hardening. To un-
strain directly by inducing back stress, so called Bauschinger effect derstand the clustering and short range ordering in high-Mn TWIP
[11,12], and indirectly through the formation of obstacles against steels, C and Mn distribution was analysed using atom probe to-
the dislocation motion, which reduces the mean free path and mography [25,34], which did not reveal any clear conclusion.
There have been several attempts to model the constitutive
enhances the dislocation multiplications, referred as the dynamic
behaviour and/or the work-hardening behaviour in austenitic
Hall-Petch effect [6–10]. Although the contribution of mechanical
metals with low SFE [8–11,23,24,29,30,37–42] and all the models
twins to the work-hardening behaviour of TWIP steels is well-
rely on the presence of the mechanical twins in the micro-
known [4,11,13–31], other microstructural phenomena on strain-
structure. In addition to the mechanical twinning, some models
ing may influence the work-hardening behaviour, such as stacking
considered the DSA effect [29] or dislocation substructure evolu-
faults (SFs) formation [27,32], clustering and short range ordering
tion [24] such as highly dense dislocation walls or dislocation cells
(SRO) of solute atoms, [25,33,34] and dynamic strain aging (DSA)
as a contributor to the enhanced work-hardening behaviour.
[4,35]. Moreover, recent study [36] using synchrotron XRD The ambiguity in the contribution of the mechanical twinning
to the work-hardening behaviour may also be related to the size of
n
Corresponding author. the mechanical twins and/or to the characterisation technique
E-mail address: hossein.beladi@deakin.edu.au (H. Beladi). approach used to analyse the twinning. Each characterisation

http://dx.doi.org/10.1016/j.msea.2016.05.104
0921-5093/& 2016 Elsevier B.V. All rights reserved.
438 V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446

method has its own advantages and limitations. For example, evolution of TWIP steel at different deformation conditions.
optical microscopy [27] could be used to characterize a large area The microstructure characterisation at different tensile testing
but could not resolve the nano-scale twins. Another method ex- conditions was performed using Zeiss Supra 55VP SEM operating at
tensively used [4,6,10,17,21,29,40] was TEM, which can resolve the 20 kV in conjunction with the angle selected backscattered (AsB)
nano-twins, however the observation area is limited. Alternative detector. AsB detector empowers to reveal nano-scale mechanical
method was scanning electron microscopy (SEM) in conjunction twinning. This enables to statistically analyse nano-scale mechan-
with EBSD detector [13,14,17,22,23,28], which can cover a larger ical twinning in a large area. The microstructure examination was
area but could not successfully reveal the twin thickness of less conducted on 26 SEM images containing a total of more than 100
than 100 nm. A newly developed Electron Channelling Contrast grains covering an area of about 300 mm2 at each deformation
Imaging (ECCI) technique [21,24,25], overcomes these limitations. condition. The volume fraction of twinned grains and the volume
However, the observations of twins appeared to be limited for the fraction of mechanical twins were measured using a point counting
grains with specific orientations [32,43,44]. technique with a statistical significance of po0.05 [45].
The current paper will meticulously investigate the micro- The work-hardening rate was estimated at all tensile testing
structure of TWIP steel, in particularly mechanical twins, formed temperatures using Eq. (1):
during deformation at different temperatures. The microstructure
∂σ
then would be correlated with the work-hardening behaviour to θ=
∂ε (1)
understand the extent of microstructural features contribution to
the work-hardening behaviour at different stage of straining. where θ is the work-hardening rate and ∂σ and ∂ε are the differ-
entials of the engineering stress and strain, respectively. The data
were smoothened every 25 points to reduce the serration.
2. Experimental procedure

The TWIP steel used in this study was provided by POSCO Steel,
3. Results
South Korea. The composition of steel was Fe-18Mn-0.6C-1Al
(wt%). The average austenite grain size was about 2.5 70.2 mm.
3.1. Mechanical properties of the samples deformed at different
The standard tensile specimens with a gauge length of 30 mm
temperatures
were wire cut from sheet product, perpendicular to the rolling
direction, with a thickness of 1.25 mm. The grip section of a tensile
The TWIP steel revealed unique mechanical properties at room
sample was modified to be wider than the standard grip area to
temperature, having a yield strength (YS) of 500 MPa, ultimate
prevent any sliding during the tensile experiment. The tensile
tensile strength (UTS) of  1000 MPa and a total elongation (TE) of
experiments were conducted at a strain rate of 10 3 s 1 using
 60% (Figs. 1 and 2). Similar uniform and total elongations in-
Instron tensile test machine with 30 kN load cell and at the tem-
dicate that the deformation was mostly uniform and ultimate
perature range of ambient ≤T [°C ] ≤ 400°C (room temperature, 100,
tensile stress occurred at a strain level close to the fracture (Fig. 2
200, 300, 400 °C). To examine the microstructural evolution dur-
(a)). The tensile testing temperature significantly affected the
ing straining, interrupted tensile tests were performed at testing
mechanical behaviour of the TWIP steel (Figs. 1 and 2). The YS and
temperatures of ambient, 100 and 200 °C.
UTS continuously decreased as the temperature increased from
Mechanical twinning is a stress-assisted mechanism. However,
ambient temperature to 400 °C. The TE showed a slight increase at
as stress and strain are interdependent, the analysis of the me-
chanical twinning as a function of strain does not change the 100 °C, followed by a significant drop at higher tensile tempera-
nature of the results. Therefore, it was easier to interrupt the tests tures of 200, 300 and 400 °C (Fig. 2(a)). Similar to the room tem-
at a given strain level and analyse the evolution of the mechanical perature condition, the uniform elongation (UE) was close to the
twinning parameters as a function of strain. total elongation at all elevated temperature testing conditions,
At ambient temperature, the tensile tests were interrupted at a suggesting that the deformation was mostly uniform at all tensile
strain of 0.04, 0.1, 0.2, 0.3, 0.4 and fracture, which was equal to temperatures (Fig. 2(a)). A decrease in both elongation and
0.5 of true strain. At 100 °C, the tensile tests were interrupted at strength with an increase in the tensile testing temperature re-
0.1 true strain intervals (i.e. 0.1, 0.2, 0.3, 0.4 and fracture, which sulted in a significant deterioration in the strength-ductility bal-
occurred at about 0.5 of true strain). At 200 °C, the tensile tests ance of the TWIP steel, represented by the UTS and total elonga-
were interrupted at a true strain of 0.1, 0.2, 0.25, 0.3, 0.35 and tion (Fig. 2(b)).
fracture, corresponding to about 0.4 of true strain. Four tensile
samples were tested for each temperature conditions. The samples 1800
used for microstructural characterisation were cut in width of Room temperature 100 C
o

10 mm from the middle of the sample or about 5 mm away from 1500


the fracture edge.
True stress (MPa)

o
The samples at room temperature were characterised using 200 C
1200
transmission electron microscopy (TEM) by Jeol 2100 with the o
300 C
voltage of 200 V. The TEM samples were sliced (2–3 slices from 900
each sample) in the longitudinal (tensile) direction and manually o
400 C
ground to a thickness of  70 mm. They were then electro-polished 600
with 5% of perchloric acid in methanol using a twin-jet Tenupol
unit, operating at 50 V and the temperature of 25 °C. Although 300
TEM is a very powerful technique to reveal the mechanical twin-
ning and dislocation substructure development, it was not possi-
0
ble to characterise the extent of twinning as a function of strain 0.0 0.1 0.2 0.3 0.4 0.5
statistically.
True strain
To overcome this problem, SEM equipped with AsB detector
was employed to statistically examine the microstructure Fig. 1. True stress-strain curves at different tensile testing temperatures.
V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446 439

Fig. 2. (a) Mechanical properties of TWIP steel (i.e. yield strength, ultimate tensile strength (UTS), total and uniform elongations) and (b) the strength-ductility balance of
TWIP steel represented by the product of UTS and total elongation (TEL), at different tensile testing temperatures.

The stress-strain curves revealed the serration for all tensile the formation of the mechanical twins in some austenite grains
testing temperatures, although the extent of serration was sig- (Fig. 6). Mechanical twin thickness was 20–50 nm and high dis-
nificantly influenced by the temperature (Fig. 1). The serration was location density was often observed in the vicinity of the me-
the most prominent at 100 °C and initiated at a true strain level of chanical twins (Fig. 6).
about 0.15, showing continuously growing amplitude with the At the strain level of 0.1, the formation of stacking faults was
deformation (Fig. 1). For other deformation temperatures, the observed homogeneously within grains (Fig. 7). Moreover, in some
serration was seen only at a very late stage of the deformation (e.g. grains stacking faults with different burger vectors were observed
a true strain level of about 0.4 and above at room temperature,
(Fig. 7(b) and (c)), which significantly increased the interaction of
Fig. 1).
the dislocations with the stacking faults (indicated by blue arrows
in Fig. 7). At this stage, stacking faults were still the predominant
3.2. Work-hardening behaviour during tensile testing at different
temperatures deformation mechanism. Moreover, the number of austenite
grains with twins increased with an increase in strain (Fig. 8). The
The work-hardening behaviour curve, i.e. dσ /dε vs. ε (Fig. 3, mechanical twins were mostly nucleated at the grain boundaries,
Table 1) of TWIP steel at room temperature can be divided into propagating across the grain interior (Fig. 8). Similar to the TEM
four stages: (I) a sharp decrease in the work-hardening until it observation, the mechanical twins revealed using AsB-SEM tech-
reached a minimum value of 2800 MPa at a true strain of about nique, were very fine having a thickness of 20–50 nm (Fig. 9(a–e).
0.025; (II) a gradual increase in the work-hardening with the de- Furthermore, some dislocation activity was observed in both
formation up to a true strain of about 0.05; (III) plateau behaviour twinned and twin-free grains. The volume fraction of twins and
from strain of 0.05 up to a true strain of about 0.07. During stage twinned grains were progressively increased with an increase in
III, the work-hardening value remained almost constant the strain (Fig. 10). The volume fraction of twinned grains and the
(  2800 MPa); and (IV) a gradual decrease with local fluctuation in volume fraction of twins enhanced from 397 6 and 0.6 70.2% at a
the overall work-hardening behaviour with straining until the
true strain level of 0.1 to 84 74 and 18.6 71.7% at a true strain
fracture occurred (Fig. 3(a)).
level of  0.5 (at fracture), respectively (Fig. 10).
The tensile testing temperature had a vital effect on the work-
The deformation temperature significantly affected the changes
hardening behaviour, influencing stages (II) and (III), (Table 1). The
in the microstructure development occurred during straining. The
stages (II) and (III) gradually diminished with an increase in the
tensile testing temperature up to 200 °C (Fig. 3(b) and (c)), above extent of the mechanical twinning was remarkably reduced with
which these stages were nearly disappeared (Fig. 3(d) and (e)). At an increase in the deformation temperature (Figs. 11 and 12), so
the tensile testing temperature of 300 and 400 °C, the work- that no mechanical twinning was observed at a deformation
hardening behaviour showed only the stages (I) and (IV) (Table 1, temperature of 300 °C and above even at fracture (Fig. 13). At this
Fig. 3d, e). temperature regime, the grains only contained dislocation sub-
structure (i.e. twin-free grains, Fig. 13). Furthermore, the initiation
3.3. Characterisation of the microstructure development at different of the mechanical twin formation was postponed from 0.04 at
deformation conditions room temperature to a true strain of 0.2 and 0.25 at 100 and
200 °C, respectively (Figs. 9(a), 11(b) and 12(b)). Mechanical twin
The TEM in the current study was used to reveal the nano-scale thickness remains almost constant throughout the deformation
twins and the dislocation substructure development at each stage regardless the deformation conditions (Figs. 9a–e, 11b–e and 12b–
of deformation. At the as-received condition, the TEM showed the
e). In general, the volume fraction of twins significantly decreased
presence of annealing twins, having a thickness of 0.2–1 mm
with an increase in the tensile deformation temperatures at a gi-
(shown by white arrows in Fig. 4). An increase in strain to 0.04 led
ven strain. At a strain level of 0.4, the volume fraction of me-
to the formation of dislocations, stacking faults (Fig. 5) and me-
chanical twins (Fig. 6). However, the stacking faults formation was chanical twins reduced from 11.2 70.7% at room temperature to
the predominant deformation mechanism at this condition. The 5 70.7 and 2.9 70.4% at 100 and 200 °C, respectively (Fig. 10(a)). A
diffraction pattern analysis confirmed the formation of stacking similar trend was observed for the volume fraction of twinned
faults, which were mostly nucleated at the grain boundaries and grains, where it was reduced from 72 75 at room temperature to
propagated into the grain interior, pinning the gliding dislocations 567 6 and 44 7 7% at 100 and 200 °C, respectively, at a strain level
(indicated by blue arrows in Fig. 5). The TEM analysis also revealed of 0.4 (Fig. 10(b)).
440 V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446

Fig. 3. Work-hardening behaviour as a function of true strain at different tensile testing temperatures, (a) at room temperature (b) 100 °C, (c) 200 °C, (d) 300 °C and
(e) 400 °C.

Table 1. 4. Discussion
Work-hardening behaviour as a function of strain at different deformation
temperatures.
The current TWIP steel reveals the exceptional mechanical
Tensile testing The stages of the work-hardening behaviour properties at room temperature, although the mechanical prop-
temperature erties are deteriorated at the elevated tensile temperatures
I II III IV
(Figs. 1 and 2). The mechanical properties are directly related to
True strain the work-hardening behaviour of the material during straining
(Fig. 3(a)). The work-hardening behaviour of the TWIP steel is
Ambient 0–0.025 0.025–0.05 0.05–0.07 0.07 onwards
highly dependent on the deformation mechanisms, which are
100 °C 0–0.018 0.018– 0.022– 0.041 onwards
0.022 0.041 associated with the stacking fault energy. Three different
200 °C 0–0.021 0.021– 0.023– 0.027 onwards deformation mechanisms are identified in austenitic materials
0.023 0.027
during straining as a function of SFE [5,23,26,42,46]: (i) slip
300 °C 0–0.024 – – 0.024 onwards
400 °C 0–0.024 – – 0.024 onwards mechanism, which is always expected to take place during
deformation, although it would be a dominant mechanism at a SFE
V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446 441

30 µm 1µm

Fig. 4. The as-received microstructure of the TWIP steel; (a) EBSD map; red and black lines represent annealing twin and high angle grain boundaries, respectively and
(b) TEM image, arrows show the annealing twins. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

002 111
131 002
[310]
[110]

0.2µm
Fig. 5. Bright field image of 0.04 strained TWIP steel showing stacking faults
emitted from the grain boundary. The inset is diffraction pattern of austenite matrix
with zone axis of [310]γ. The red arrows and dash line represent the stacking faults
and the grain boundary, respectively. The interactions of the glide dislocations with
the stacking faults are shown by blue arrows. (For interpretation of the references
0.5 µm
to color in this figure legend, the reader is referred to the web version of this
article.) Fig. 6. Bright field image of 0.04 strained TWIP steel showing deformation twins.
The inset is diffraction pattern of mechanically twinned austenite with zone axis of
[110]γ. Red arrow represents the diffraction spot from the mechanical twins. (For
greater than 60 mJ/m2 [5,23,26,42,46]; (ii) transformation of aus- interpretation of the references to color in this figure legend, the reader is referred
tenite to martensite, Transformation Induced Plasticity (TRIP) ef- to the web version of this article.)
fect, in austenitic steels with a SFE lower than 20 mJ/m2
[5,23,26,42,46]; and (iii) twinning of austenite (TWIP effect) in in most materials regardless of their SFE (Fig. 15). This can largely
austenitic steels with a SFE between 20–60 mJ/m2 [5,23,26,42,46]. be related to the elastic strain and transition from elastic to plastic
The last two deformation mechanisms result in the formation of deformation [51]. There is a minima at the end of stage (I) (i.e.
new obstacles against dislocations during deformation, which  0.025 strain) followed by an increase in the work-hardening rate
enhances the work-hardening rate. As a result, necking is delayed, up to a strain of  0.05 (i.e. stage (II), Figs. 3(a) and 15). The work-
which improves the plasticity of the material. At room tempera- hardening behaviour can be associated with different micro-
ture, the SFE of current TWIP steel is measured in a range of structural features such as mechanical twinning [20–22,26], grain
14–29 mJ/m2 using different thermodynamic approaches boundaries [52], stacking faults formation [27,32], dislocation
[23,42,44,47] (Fig. 14), which is in a good agreement with the SFE cross slip [53] and/or DSA [4,35]. The mechanical twinning is be-
results measured by TEM [48] and XRD [49] for an austenitic steel lieved to significantly influence the work-hardening as the initia-
with a similar composition. In addition, the martensitic transfor- tion of mechanical twins continuously limits the extent of the
mation does not occur in the similar steel composition confirmed dislocation glide by blocking their movements [20–22,26]. Indeed,
by others [50] using EBSD and XRD techniques. the mechanical twins propagate across the grains, progressively
The work-hardening behaviour at room temperature consisted reducing the dislocation mean free path during straining and in-
of four distinct stages, which could be closely associated with the troducing the dynamic Hall-Petch effect. In the current study, the
microstructural changes during deformation (Fig. 3(a)). This was mechanical twinning is, however, observed at a strain level of
demonstrated schematically in Fig. 15. The first stage reveals a  0.04, which is just below the maximum strain associated with
typical intense reduction in the work-hardening rate, as observed the end of the stage (II) of work-hardening, i.e. 0.05 (Figs. 3(a) and
442 V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446

111 002
b 220
[110] 111
[112]

200 nm 100 nm 100 nm

Fig. 7. (a–b) Bright field images of stacking faults formed after 0.1 strain. The insets in (a) and (b) are diffraction patterns of austenite with a zone axis of [110]γ and [112]γ,
respectively. (c) the dark field image of (b) from (11-1)γ reflection. The red and blue arrows represent the stacking faults and the interactions of the glide dislocations with
the stacking faults, respectively. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

15). In addition, the volume fraction of mechanical twins is very This suggests that the elevated tensile testing temperature sig-
small, i.e.  0.02% (Fig. 10(a)), at this strain range. This suggests nificantly influences the interactions of glide dislocations with
that an increase observed at stage (II) of the work-hardening is less stacking faults and grain boundaries. The stacking fault energy
likely attributed to the mechanical twinning at room temperature. increased with temperature (Fig. 14), enhancing the separation
The DSA can be also rolled out here as it takes place beyond a true distance of dissociated partial dislocations, which consequently
strain of 0.4 at room temperature, which will be discussed in more reduces the interactions of glide dislocations with stacking faults.
details further on. The current steel has a small grain size of Similarly, the pinning of dislocations against grain boundaries
2.570.2 mm, having a relatively high grain boundary area, which progressively becomes less effective with temperature. This is
significantly enhances the chance of interactions between the dis- more obvious at the temperature of 300 °C and above, where stage
locations and the grain boundaries [52]. These interactions can (II) of the work-hardening behaviour almost disappears (Fig. 3
partly contribute to the augmentation of the work-hardening rate (d) and (e)). At this temperature regime, the stacking fault energy
[52]. Another reason for an increase in the work-hardening rate at is estimated beyond 60 mJ/m2, proposing that the slip is a pre-
an early stage of deformation can be attributed to the formation of dominant deformation mechanism (i.e. no mechanical twin for-
the stacking faults [32]. In the current study, the stacking faults are mation) as supported by the current result.
frequently observed in the vicinity of grain boundaries at a strain of Stage (III) of the work-hardening behaviour is referred to the
0.04 at room temperature (Fig. 5). At strain level of 0.1, stacking plateau region observed at a strain range of 0.05–0.07 at room
faults are, however, homogeneously spread inside the grains and temperature (Figs. 3(a) and 15). This could be referred to the
the gliding dislocations extensively interact with them (Fig. 7). The stacking faults and/or the formation of mechanical twins. Inter-
stacking faults are formed as a result of the dissociation of the estingly, the volume fraction of mechanical twinning is less than
perfect dislocations into Shockley partials in materials with low SFE 0.6% at this strain range (Fig. 10(a)), suggesting that the interac-
[27,32]. They can be widely extended and act as a source of dis- tions of glide dislocations with stacking faults have greater con-
locations production and their interaction with glide dislocations is tributions to the stage (III) of the work-hardening behaviour
known to increase the work-hardening rate [32]. compared with the mechanical twinning. The temperature rise
Interestingly, the stage (II) of the work-hardening behaviour significantly reduces the strain range of stage (III), and it is almost
progressively becomes narrower with the temperature (Fig. 3). vanished at a temperature above 200 °C (Fig. 3(d) and (e)). The SFE

2µm 2µm
Fig. 8. Representative TEM image of austenite microstructure at different true strain levels of (a) 0.2, (b) 0.4 at room temperature. White arrows indicate the mechanical
twins.
V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446 443

500 nm 500 nm 500 nm

500 nm 500 nm

Fig. 9. SEM images of the microstructure at different true strain levels of (a) 0.1, (b) 0.2, (c) 0.3, (d) 0.4 and (e) fracture (  0.5) at room temperature.

Volume fraction of twinned grains (%)


100
RT
20
Volume fraction of twins (%)

o
100 C
o 80
200 C
15
60

10
40

5 RT
20 o
100 C
o
200 C
0 0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
True strain True strain
Fig. 10. (a) Volume fraction of mechanical twins and (b) the dislocation mean free path of twinned grains as a function of true strain at different tensile testing temperatures.

500 nm 500 nm 500 nm

500 nm 500 nm

Fig. 11. SEM images of the microstructure at different true strain levels of (a) 0.1, (b) 0.2, (c) 0.3, (d) 0.4 and (e) fracture (  0.5) at 100 °C.
444 V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446

500 nm 500 nm 500 nm

500 nm 500 nm

Fig. 12. SEM images of the microstructure at different true strain levels of (a) 0.2, (b) 0.25, (c) 0.3, (d) 0.35 and (e) fracture (  0.4) at 200 °C.

500 nm 500 nm

Fig. 13. SEM images of the microstructure at fracture at different tensile testing temperatures: (a) 300 °C and (b) 400 °C.

120
Work-hardening (d /d )

Allain 2004 DG GBs SFs and SFs, Twins and


Dumay et al. 2007 and SFs Twins DSA
100 Mazancova and Mazanec 2009 (i) (ii) (iii) (iv)
A.Saeed-Akbari 2009
80
SFE (mJ/m )

Slip
2

60 Low SFE
TWIP steel with fine grains

40 TWIP High SFE material


with fine grains

20
Medium to high SFE material
TRIP with coarse grains
0
0 50 100 150 200 250 300 350 400 True strain
o
Temperature ( C) Fig. 15. Schematic representation of the work-hardening behaviour of fcc materials
with different stacking fault energies (SFE) under tensile testing. Dash line curve
Fig. 14. Calculated stacking fault energy of the TWIP steel as a function of tem- represents the material with medium to high SFE having coarse grain size. Dot line
perature using different thermodynamic approaches. curve shows the material with medium to high SFE with small grain size. Solid line
curve represents the low SFE TWIP steel with fine grain. DG, GB, SF and DSA are
dislocation glide, grain boundaries, stacking faults and dynamic strain aging,
at a temperature above 200 °C enhances to such extent that sig- respectively.
nificantly reduces the effectiveness of the dislocation and stacking
faults interaction and postpones the formation of mechanical what is observed in stage (I). Interestingly, the current result re-
twins. veals a progressive augmentation in the mechanical twinning
Beyond a strain of 0.07, the work-hardening rate gradually volume fraction beyond a strain of 0.04 at room temperature
decreases at room temperature (i.e. stage (IV), Figs. 3(a) and 15), (Fig. 10(a)). It would be expected that the mechanical twinning
though the rate of work-hardening decline is much slower than enhances the work-hardening rate through dynamic Hall-Petch
V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446 445

effect. This suggests that the dislocation annihilation (i.e. recovery) Industry, InTech, 2011.
rate is still greater than the dislocation multiplication (i.e. hard- [2] B.C. De Cooman, et al., State-of-the-Science of High Manganese TWIP Steels for
Automotive Applications, Springer, London, 2009.
ening) rate induced by the mechanical twinning and/or stacking [3] O. Bouaziz, et al., High manganese austenitic twinning induced plasticity
faults. At high strain levels, i.e. beyond 0.35 at room temperature, steels: A review of the microstructure properties relationships, Curr. Opin.
the work-hardening behaviour reveals a significant oscillation as Solid State Mater. Sci. 15 (4) (2011) 141–168.
[4] T. Shun, C.M. Wan, J.G. Byrne, A study of work hardening in austenitic Fe—
their amplitude augmented with strain. This behaviour can be Mn—C and Fe—Mn—Al—C alloys, Acta Metall. Mater. 40 (12) (1992) 3407–3412.
referred to DSA effect [43,54], which is commonly observed in [5] L. Remy, A. Pineau, Twinning and Strain-Induced F.C.C. - H.C.P. Transforma-
high-Mn austenitic steels and is more evident in alloys with a C tion in the FeMnCrC System, Mater. Sci. Eng. 28 (1) (1977) 99–107.
[6] O. Grässel, et al., High strength Fe-Mn-(Al, Si) TRIP/TWIP steels development –
concentration of more than 0.5 wt% [43]. The DSA effect is mainly properties – application, Int. J. Plast. 16 (10–11) (2000) 1391–1409.
referred to the interaction of C atoms of the C-Mn bundle with the [7] I. Karaman, et al., Deformation of single crystal Hadfield steel by twinning and
stacking faults during straining [54], although its contribution to slip, Acta Mater. 48 (6) (2000) 1345–1359.
[8] O. Bouaziz, N. Guelton, Modelling of TWIP effect on work-hardening, Mater.
work-hardening is relatively low at room temperature (  20 MPa)
Sci. Eng.: A 319–321 (2001) 246–249.
[55,56]. It is worth mentioning that the initiation of DSA takes [9] S. Allain, J.P. Chateau, O. Bouaziz, A physical model of the twinning-induced
place at a relatively late stage of straining at room temperature. plasticity effect in a high manganese austenitic steel, Mater. Sci. Eng.: A 387–
389 (1–2) (2004) 143–147.
This is most likely due to the presence of Al in composition, which
[10] S. Allain, et al., Modeling of mechanical twinning in a high manganese content
significantly delays the DSA initiation [50]. austenitic steel, Mater. Sci. Eng.: A 387–389 (1–2) (2004) 272–276.
Interestingly, the DSA is remarkably enhanced at a tensile [11] O. Bouaziz, S. Allain, C. Scott, Effect of grain and twin boundaries on the
temperature of 100 °C and hardly apparent at a temperature of hardening mechanisms of twinning-induced plasticity steels, Scr. Mater. 58 (6)
(2008) 484–487.
200 °C and above (Fig. 1). The DSA is significantly affected by steel [12] M. Huang, et al., Modelling the effect of carbon on deformation behaviour of
composition and tends to operate in a specific temperature range twinning induced plasticity steels, J. Mater. Sci. (2011) 1–5.
[4,35,57]. This suggests that Al shifts the DSA temperature in a [13] D. Barbier, et al., EBSD for analysing the twinning microstructure in fine-
grained TWIP steels and its influence on work hardening, J. Microsc. 235 (1)
range of room temperature to 100 °C. (2009) 67–78.
[14] J.-E. Jin, Y.-K. Lee, Strain hardening behavior of a Fe-18Mn-0.6C-1.5Al TWIP
steel, Mater. Sci. Eng.: A 527 (1–2) (2009) 157–161.
[15] R. Ueji, et al., Effect of Grain Size on the work-hardening behavior and de-
5. Summary formation twinning of high-manganese austenitic steel, Mater. Charact. (2009)
4.
In the current study, the effect of strain on the TWIP steel [16] C. Efstathiou, H. Sehitoglu, Strain hardening and heterogeneous deformation
during twinning in Hadfield steel, Acta Mater. 58 (5) (2010) 1479–1488.
microstructure development at different deformation tempera- [17] H. Beladi, et al., Orientation dependence of twinning and strain hardening
tures was analysed. The microstructural changes were correlated behaviour of a high manganese twinning induced plasticity steel with poly-
with the work-hardening behaviour of TWIP steel at different crystalline structure, Acta Mater. 59 (20) (2011) 7787–7799.
[18] O. Bouaziz, et al., Effect of chemical composition on work hardening of Fe-Mn-
deformation temperatures. The important findings of present C TWIP steels, Mater. Sci. Technol. 27 (3) (2011) 707–709.
study were: [19] H. Ding, et al., Strain hardening behavior of a TRIP/TWIP steel with 18.8% Mn,
Mater. Sci. Eng.: A 528 (3) (2011) 868–873.
[20] I. Gutierrez-Urrutia, D. Raabe, Dislocation and twin substructure evolution
(1) the current TWIP steel revealed an exceptional mechanical
during strain hardening of an Fe–22 wt% Mn–0.6 wt% C TWIP steel observed
behaviour at room temperature, which gradually deteriorated by electron channeling contrast imaging, Acta Mater. 59 (16) (2011)
with an increase in the tensile testing temperature. 6449–6462.
[21] I. Gutierrez-Urrutia, D. Raabe, Multistage strain hardening through dislocation
(2) the work-hardening behaviour revealed four distinct stages at
substructure and twinning in a high strength and ductile weight-reduced Fe–
room temperature, though the extent of these stages was Mn–Al–C steel, Acta Mater. 60 (16) (2012) 5791–5802.
significantly influenced by the tensile testing temperature. [22] K. Renard, P.J. Jacques, On the relationship between work hardening and
(3) the work-hardening behaviour was attributed to complex twinning rate in TWIP steels, Mater. Sci. Eng.: A 542 (0) (2012) 8–14.
[23] A. Saeed-Akbari, et al., Characterization and prediction of flow behavior in
dynamic strain induced microstructural reactions consisted of high-manganese twinning induced plasticity steels: Part I. Mechanism maps
dynamic recovery, dislocation dissociation, stacking fault for- and work-hardening behavior, Metall. Mater. Trans. A 43 (5) (2012)
mation, mechanical twining and dynamic strain aging. 1688–1704.
[24] D.R. Steinmetz, et al., Revealing the strain-hardening behavior of twinning-
(4) the current observation suggested that the mechanical twins induced plasticity steels: theory, simulations, experiments, Acta Mater. 61 (2)
had a limited effect on the work-hardening behaviour of the (2013) 494–510.
TWIP steel and its influence significantly reduced with the [25] I. Gutierrez-Urrutia et al., Revealing the strain-hardening mechanisms of ad-
vanced high-Mn steels by multi-scale microstructure characterization, in:
temperature. The development of stacking faults and their
Proceedings of the 8th International Conference on Processing and Manu-
interactions with the glide dislocations is believed to be the facturing of Advanced Materials, THERMEC 2013, 2014, Trans Tech Publica-
main contributor in the work-hardening enhancement of the tions Ltd: Las Vegas, NV, pp. 755–760.
TWIP steel. [26] V. Shterner, I.B. Timokhina, H. Beladi, The correlation between stacking fault
energy and the work hardening behaviour of high-mn twinning induced
plasticity steel tested at various temperatures, in: Proceedings of the 8th In-
ternational Conference on Processing and Manufacturing of Advanced Mate-
Acknowledgement rials: Processing, Fabrication, Properties, Applications, THERMEC 2013, 2014,
Trans Tech Publications: Las Vegas, NV, pp. 676–681.
[27] L. Rémy, The interaction between slip and twinning systems and the influence
Prof. B.C. De Cooman is thankfully acknowledged for providing of twinning on the mechanical behavior of fcc metals and alloys, Metall. Mater.
the TWIP steel. This study was supported by grants provided by Trans. A 12 (3) (1981) 387–408.
[28] D. Barbier, et al., Analysis of the tensile behavior of a TWIP steel based on the
the Australian Research Council (Grant no. FL0992361). The au- texture and microstructure evolutions, Mater. Sci. Eng. A 500 (1–2) (2009)
thors would like to acknowledge the technical and scientific sup- 196–206.
port of the Deakin Advanced Characterisation Facility and Monash [29] J. Kim, et al., Constitutive modeling of the tensile behavior of Al-TWIP steel,
Metall. Mater. Trans. A 43 (2) (2012) 479–490.
Centre for Electron Microscopy. VS also acknowledges Deakin [30] V. Shterner, et al., A constitutive model of the deformation behaviour of
University for providing a research scholarship for this work. twinning induced plasticity (TWIP) steel at different temperatures, Mater. Sci.
Eng.: A 613 (0) (2014) 224–231.
References [31] V. Shterner, et al., Effect of Temperature on Mechanical Behaviour of High
Manganese TWIP steel, Wollongong, NSW 2014, pp. 257–262.
[32] D. Drobnjak, J. Parr, Deformation substructure and strain-hardening char-
[1] B. De Cooman, K.-g. Chin, J. Kim, High Mn TWIP steels for automotive appli- acteristics of metastable Fe-Mn austenites, Metall. Mater. Trans. B 1 (4) (1970)
cations, in: M. Chiaberge (Ed.), New Trends and Developments in Automotive 759–765.
446 V. Shterner et al. / Materials Science & Engineering A 669 (2016) 437–446

[33] F. Hamdi, S. Asgari, Evaluation of the role of deformation twinning in work [45] J.C. Russ, R.T. Dehoff, Practical Stereology, 2nd ed., Plenum Press, New York, NY
hardening behavior of face-centered-cubic polycrystals, Metall. Mater. Trans. A 1999, p. 312.
39 (2) (2008) 294–303. [46] A. Saeed-Akbari, et al., Derivation and variation in composition-dependent
[34] R.K.W. Marceau, P. Choi, D. Raabe, Understanding the detection of carbon in stacking fault energy maps based on subregular solution model in high-
austenitic high-Mn steel using atom probe tomography, Ultramicroscopy 132 manganese steels, Metall. Mater. Trans. A 40 (13) (2009) 3076–3090.
(0) (2013) 239–247. [47] A. Dumay, et al., Influence of addition elements on the stacking-fault energy
[35] Y. Dastur, W. Leslie, Mechanism of work hardening in Hadfield manganese and mechanical properties of an austenitic Fe–Mn–C steel, Mater. Sci. Eng.: A
steel, Metall. Mater. Trans. A 12 (5) (1981) 749–759. 483–484 (0) (2008) 184–187.
[36] Z.Y. Liang, Y.Z. Li, M.X. Huang, The respective hardening contributions of dis- [48] J. Kim, B. De Cooman, On the stacking fault energy of Fe-18 Pct Mn-0.6 Pct
locations and twins to the flow stress of a twinning-induced plasticity steel, C-1.5 Pct Al twinning-induced plasticity steel, Metall. Mater. Trans. A 42 (4)
Scr. Mater. 112 (2016) 28–31. (2011) 932–936.
[37] J. Gil Sevillano, P. van Houtte, E. Aernoudt, Large strain work hardening and [49] J.S. Jeong, et al., In situ neutron diffraction study of the microstructure and
textures, Prog. Mater. Sci. 25 (2) (1980) 69–134. tensile deformation behavior in Al-added high manganese austenitic steels,
[38] H. Mecking, U.F. Kocks, Kinetics of flow and strain-hardening, Acta Metall. 29 Acta Mater. 60 (5) (2012) 2290–2299.
(11) (1981) 1865–1875. [50] J.-K. Kim, et al., On the tensile behavior of high-manganese twinning-induced
[39] S.R. Kalidindi, Modeling the strain hardening response of low SFE FCC alloys, plasticity steel, Metall. Mater. Trans. A 40 (13) (2009) 3147–3158.
Int. J. Plast. 14 (12) (1998) 1265–1277. [51] U.F. Kocks, H. Mecking, Physics and phenomenology of strain hardening: the
[40] I. Karaman, et al., Competing mechanisms and modeling of deformation in FCC case, Prog. Mater. Sci. 48 (3) (2003) 171–273.
austenitic stainless steel single crystals with and without nitrogen, Acta Mater. [52] C.W. Sinclair, W.J. Poole, Y. Bréchet, A model for the grain size dependent work
49 (19) (2001) 3919–3933. hardening of copper, Scr. Mater. 55 (8) (2006) 739–742.
[41] J. Gil Sevillano, An alternative model for the strain hardening of FCC alloys that [53] W. Püschl, Models for dislocation cross-slip in close-packed crystal structures:
twin, validated for twinning-induced plasticity steel, Scr. Mater. 60 (5) (2009) a critical review, Prog. Mater. Sci. 47 (4) (2002) 415–461.
336–339. [54] S.-J. Lee, et al., On the origin of dynamic strain aging in twinning-induced
[42] S. Allain, et al., Correlations between the calculated stacking fault energy and plasticity steels, Acta Mater. 59 (17) (2011) 6809–6819.
the plasticity mechanisms in Fe–Mn–C alloys, Mater. Sci. Eng.: A 387–389 (0) [55] R.A. Mulford, U.F. Kocks, New observations on the mechanisms of dynamic
(2004) 158–162. strain aging and of jerky flow, Acta Metall. 27 (7) (1979) 1125–1134.
[43] L. Chen, et al., Localized deformation due to Portevin-LeChatelier effect in [56] A. Wijler, M.M.A. Vrijhoef, A. Van Den Beukel, The onset of serrated yielding in
18Mn-0.6C TWIP austenitic steel, ISIJ Int. 47 (12) (2007) 1804–1812. Au(Cu) alloys, Acta Metall. 22 (1) (1974) 13–19.
[44] E. Mazancova, K. Mazanec, Stacking fault energy in high manganese alloys, [57] D. Drobnjak, J.G. Parr, Deformation substructure and strain-hardening char-
Mater. Eng. 16 (2) (2009) 6. acteristics of metastable Fe-Mn austenites, Metall. Trans. 1 (4) (1970) 759–765.

You might also like