Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Science & Engineering A 767 (2019) 138406

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

The effect of reduced temperatures on microstructure development in tensile T


tested high-manganese steel
Joanna Kowalskaa,∗, Janusz Ryśa, Grzegorz Ciosb, Wiktor Bednarczyka
a
AGH University of Science and Technology, Faculty of Metals Engineering and Industrial Computer Science, Cracow, Poland
b
AGH University of Science and Technology, Academic Centre for Materials and Nanotechnology, Cracow, Poland

A R T I C LE I N FO A B S T R A C T

Keywords: The current research is part of a project covering several grades of high-manganese steels (20–26 wt% Mn) with
High-manganese steel different chemical compositions, which were deformed at room and reduced temperatures. In the present paper,
Stacking fault energy the microstructure development is analyzed at the onset of plastic deformation in Fe–Mn–Al–Si steel containing
Tensile deformation 21.2 wt% Mn. The examination was conducted under conditions of lowering the deformation temperature and
Microstructure
aimed to show the influence of reduced temperatures on the occurrence of strain induced phase (martensitic)
Phase transformation
Martensite
transformation.
The steel samples were subjected to uniaxial deformation in a tensile test at gradually reduced temperatures,
starting from 20 °C and then applying the temperatures 0 °C, −50 °C and −150 °C. The present paper includes
the analysis of microstructure development and orientation relationships as well as determination of operating
deformation mechanisms for samples after 10% of tensile strain. The main part of the research was carried out by
means of scanning electron microscopy (SEM), using the method of electron back-scatter diffraction (EBSD).
The carried out analysis indicates that within the range of small strains the mechanism controlling de-
formation behavior of high-manganese steel Fe–21Mn–3Al–3Si was slip of partial dislocations. This was im-
mediately followed by the occurrence of strain induced phase transformation γ (fcc) → ε (hcp) and then locally ε
(hcp) → α' (bcc). With decreasing deformation temperature a noticeable increase of the volume fraction of the
martensitic phases was observed, first of all the ε-phase and to a much lesser extent α′-martensite. The ε-mar-
tensite appeared on the background of a deformed austenite initially in the form of relatively narrow bands. The
α′-martensite was formed predominantly at the intersections of the ε-bands. An interesting result, however, is
the fact that numerous places were also found where α′-martensite was formed in single set of parallel ε-phase
bands. The preferential crystallographic orientation relations (ORs) between the γ-austenite and the strain in-
duced ε- and α′- phases were successively followed, showing first of all the appearance of Shoji-Nishiyama (S–N)
and Kurdjumov-Sachs (K–S) ORs respectively.

1. Introduction heavy loads. High-manganese steels show an advantageous combina-


tion of strength and plastic properties which is related to the mechan-
Austenitic steels with a higher manganese contents have attracted isms of plastic deformation. Depending on chemical composition and
attention and remained at the centre of interest practically since the end temperature, they exhibit different values of stacking fault energy (SFE)
of the 19th century. Discovered by Hadfield in 1882, manganese steels which strongly influences deformation mechanisms, and as a result
are regarded as one of the first alloy steels. These materials show ra- final mechanical properties [1,3,7–11]. Together with changes in the
dically higher mechanical properties in comparison to plain carbon SFE value, the required toughness and strength are usually achieved by
steels and first of all have no equal in their ability to work-harden and the combined operation of three different deformation mechanisms.
to absorb impact energy. First of all this is the mechanism of dislocation slip and more specifi-
Over the last few decades, an increasing scientific and technological cally, the slip of Shockley partials separated by stacking faults. The
interest in high-manganese steels was observed [1–6]. Currently, these second mechanism is mechanical twinning, which should increase the
materials are the next generation of steels for the automotive structural total elongation and ensure the pronounced hardening rate. While the
parts as well as a number of other structural applications exposed to third one, is the strain-induced phase transformation, resulting in the


Corresponding author.
E-mail address: joannak@agh.edu.pl (J. Kowalska).

https://doi.org/10.1016/j.msea.2019.138406
Received 25 July 2019; Received in revised form 28 August 2019; Accepted 8 September 2019
Available online 09 September 2019
0921-5093/ © 2019 Elsevier B.V. All rights reserved.
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

appearance of the ε- and/or α′-martensite within the microstructure. deformation mechanism should be twinning.
This mechanism is supposed to increase the yield strength and hard- Most of the investigations conducted so far on high-manganese
ening rate of the steel. steels concerned the influence of higher or elevated temperatures on
Taking the above into account, the high-manganese steels are called microstructure development and consequently the resulting TWIP effect
steels with the TWIP effect, if the mechanical twinning is the me- (e.g. Refs. [19–26]). On the other hand, much less research was devoted
chanism controlling plastic flow (TWIP - twinning induced plasticity) or to the plastic deformation of austenitic steels at lower temperatures
steels with the TRIP effect, if the martensitic phase transformations [4,17,27]. That is why, the purpose of this study was to examine the
occur during deformation (TRIP - transformation induced plasticity) influence of reduced temperatures on deformation behaviour, micro-
[7–11]. In general, both effects improve work hardening rate and delay structure development and mechanical properties of high-manganese
the necking before fracture, resulting in a very good combination of steel containing 21.2 wt% Mn. The focus was on the initial stage of
strength and ductility. It should be emphasized however, that the plastic deformation, because the very beginning of the process clearly
analysis of strain induced phase transformations usually appears a indicated at structural instability of austenite and thus the γ (fcc) → ε
complex problem because of a diversity of nucleation sites, formation (hcp) phase transformation as a mechanisms controlling deformation.
mechanisms and transformation sequences, including the most fre- Attention was also paid to the formation mechanism and nucleation
quently reported: γ (fcc) → ε (hcp) → α' (bcc) or γ (fcc) → twin → α' sites of α′-martensite in the initial stages of deformation.
(bcc) as well as direct γ (fcc) → α' (bcc).
Numerous hitherto studies indicate that SFE value is the basic factor
determining the occurrence of mechanisms that control the plastic flow 2. Material and experimental procedure
of high-manganese steels. At this point it should be noted, that in low
stacking fault energy fcc alloys, perfect (or unit) dislocations are dis- The starting material for the experiment was high-manganese steel
sociated into two Shockley partials, which are separated by ribbons of Fe–Mn–Al–Si after hot-rolling and subsequent solution treatment. The
stacking faults. Hence, the appearance of SFs results from the passage of model steel was obtained by laboratory melting in vacuum induction
partial dislocations with Burgers vectors a/6 < 112 > on {111} slip furnace VSG-100PVA TePla AG. Casting was carried out in argon at-
planes and in consequence the sequence of close packed planes across mosphere. Analysis of the chemical composition of the steel (given in
the associated stacking fault becomes effectively hcp [12]. The forma- Table 1) was determined by the ICP-OES method (Inductively Coupled
tion of the ε-phase and twins results from different spatial arrangement Plasma - Optical Emission Spectroscopy) using Ultima 2 spectrometer
of stacking faults, i.e. the character of their overlapping. The ε-phase is by Jobin-Yvon. After casting, steel homogenization at the temperature
formed if overlapping of SFs proceeds on alternate {111} planes (i.e. 1150 °C and then hot-rolling were conducted. Finally the steel was
every second close-packed plane), while deformation twins are pro- annealed in argon atmosphere at the same temperature 1150 °C for 1 h
duced when SFs overlap on successive {111} planes (e.g. Refs. followed by water quenching.
[13–15]). Steel samples suitable for tensile testing were cut from the material
According to Dumay et al. [16] for the SFE values above 18 mJ/m2 after the preliminary treatment. Tensile tests were carried out at four
the alloys show tendency to mechanical twinning, whereas for smaller temperatures, i.e. 20 °C, 0 °C, −50 °C and −150 °C, on test gauge bars
SFE values the TRIP effect dominates upon plastic deformation of these having a diameter of 5.0 mm and a gauge length of 50.0 mm. The
steels. Also Grässel et al. [1] published the work concerning the TRIP/ tensile machine used was an Instron 5982 apparatus and the applied
TWIP effect in high-manganese steels. They examined the influence of strain rate was 10−3s−1. For the examination presented in this paper,
alloying elements on mechanical properties of the austenitic steels Fe the material was taken in the initial state (0%) and after a deformation
(15–30 wt%) Mn with Al and Si additions. The authors found that alloys of 10% for each of the selected test temperatures. It is to note, that
show the TRIP effect if the content of Mn was below 20 wt%, whereas tensile curves were not corrected from view point of machine com-
the effect of TWIP dominated in the case of manganese content higher pliance and each temperature variant was performed to sample fracture
than 25 wt% [1]. On the other hand Bouaziz et al. [3] found that if the and carefully up to 10%.
content of manganese is between 15 and 25 wt%, the structural effects X-ray phase analysis was carried out on a Siemens D500 dif-
of both types of transformations co-exist within the microstructure of fractometer using a lamp with a copper anode. Diffraction analysis was
the steel after deformation. Addition of silicon into the high-manganese performed on a longitudinal-section, parallel to the direction of
steel reduces SFE and favors the γ (fcc) → ε (hcp) transformation, both straining, in a middle part of the gauge length. The samples were pre-
during cooling and deformation [7]. On the contrary, it was shown that viously ground on abrasive papers, polished on diamond pastes and
aluminum causes an increase in SFE value, and therefore strongly electro-polished to remove the deformed layer.
suppresses the γ (fcc) → ε (hcp) transformation [8,17]. Increasing the Samples for the metallographic examination were also taken from a
carbon content in high-manganese steel raises SFE and stabilizes the longitudinal section parallel to the direction of straining.
austenite. Ghasri-Khouzani and McDermid [9] investigated the effect of Microstructure observations and analysis were performed using the FEI
carbon content on the properties and deformation structure of the QUANTA 3D FEGSEM scanning electron microscope. A method of
Fe–22Mn–C steel. During deformation of the steel with a lower carbon backscattered electron diffraction (EBSD) was applied, using EDAX
content, the transformation of γ-austenite into ε-martensite dominated, Hikari camera with the following system parameters: beam current
whereas the mechanical twinning controlled plastic flow in the alloy 20 nA, acceleration voltage 20 kV, binning 4 × 4 (120 × 120 effective
with a higher carbon content, about 0.6 wt % C. In turn, Shen et al. [18] pixel resolution), step size of 100 nm and acquisition time of approxi-
examined the steel Fe–20Mn-0,6C with similar chemical composition, mately 5 ms.
which also showed the TWIP effect during deformation. However, the
essence of the study was the analysis of the influence of grain size and
strain rate on the course of twinning process and the resulting combi-
nation of plastic and strength properties [18].
A well-known experimental result is that the temperature sig- Table 1
nificantly influences the SFE value of alloys and in consequence affects The chemical composition of high-manganese steel under examination, in wt.%.
mechanisms operating upon deformation [11]. Lowering the tempera-
Mn Si Al C P S Fe
ture leads to a decrease in SFE and thus martensite should form during
plastic deformation of austenitic steels. In turn, the increase of the 21.2 2.99 2.73 0.02 0.005 0.014 balance
deformation temperature results in the increase of SFE, so the dominant

2
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

initial state (0%) and after the tensile strain of 10% at different test
temperatures are shown in Fig. 2a and b. The diffraction pattern of the
starting material (0%) shows only lines originating essentially from
austenite, first of all (111)γ and (200)γ, as well as (220)γ, (311)γ and
(222)γ. It should be emphasized however, that the small contribution of
the α-ferrite was also detected after the preliminary treatment, reaching
roughly up to about 3% of volume fraction. After 10% of tensile strain,
clear and strong peaks appeared on diffraction patterns coming from
the ε-martensite having the hexagonal structure, namely the (10‾ 1 0) ε
and (10‾ 1 1) ε, as well as the weaker ones (10‾1 2) ε and (20‾
2 1) ε. These
peaks began to appear at reduced temperatures, starting from 0 °C. A
similar result was obtained by Koyama et al. [28] in the case of cold
deformed high-manganese steel Fe–17Mn-0.6C with increased carbon
content. The authors also observed the phase transformation induced
by strain γ (fcc) → ε (hcp) at deformation temperature of 0 °C, although
Fig. 1. Stress-strain curves for the Fe-Mn-Al-Si steel samples tensile tested at at ambient temperature of 20 °C the austenite remained essentially a
reduced temperatures. stable phase.
In addition to the initial α-ferrite, clear diffraction lines coming
3. Results and discussion from the strain-induced bcc phase were also detected in deformed
material, it means peaks from the α′-martensite. In general the intensity
3.1. Mechanical testing of the martensite diffraction lines (ε and α′) increased with decreasing
deformation temperature. For deformation temperature of −150 °C, a
Stress-strain curves for the Fe–Mn–Al–Si steel under examination, certain decrease in the intensity of the lines originating from the ε-
obtained at reduced deformation temperatures, are shown in Fig. 1. The phase at the expense of the α′-martensite can be observed. It should be
results regarding ultimate tensile strength, total elongation and ad- noted however, that the X-ray phase analysis of the steel under ex-
ditionally Vickers micro-hardness, are juxtaposed for each test tem- amination faces certain difficulties, because some diffraction lines
perature in Table 2. It is evident that the values of ultimate tensile coming from the analyzed phases overlap for given values of 2θ angle,
strength increase significantly as the temperature decreases, starting for example (111)γ and (0002)ε, (220)γ and (11‾ 2 0)ε as well as (222)γ
from 680 MPa at the temperature 20 °C up to 845 MPa and 1017 MPa at and (0004)ε.
−50 °C and −150 °C respectively. Simultaneously a parameter char- Very significant influence of temperature on the occurrence of strain
acterizing plastic properties, i.e. the total elongation, decreases its value induced martensitic transformation and twining was also found in the
from 65% at the temperature 0 °C to 58% and 54% at −50 °C and research of Fang et al. [27]. The authors examined high manganese
−150 °C respectively. It should be noted, that the values of both (26 wt% Mn) steel Fe–Mn–Al–Si with low carbon content, tensile tested
parameters remain very high for each of the test temperatures (Table 2) within the temperature range 500 °C to −70 °C. On the diffractogram of
compared to plain carbon steels. This comparison indicates at a very the sample in the initial state only the austenite was identified, whereas
good combination of plastic and strength properties of investigated after deformation at 20 °C and −70 °C the austenite turned out to be a
steel. Similarly, a value of another quantity that characterizes strength metastable phase during deformation and partially transformed into the
properties, i.e. Vickers micro-hardness, also increases markedly with a ε-martensite. Also in this case the X-ray phase analysis clearly in-
decrease in test temperature (244–294 HV0.1). dicated, that the reduction of deformation temperature favored a par-
The microstructure analysis undertaken in this paper concerns tial phase transformation γ (fcc) → ε (hcp) [27].
samples deformed up to 10% of tensile strain. It is necessary therefore,
to compare the rate of work hardening at different test temperatures for 3.3. Microstructure examination
this level of deformation. The stress-strain curves in Fig. 1 indicate that,
in the range of small deformations, differences in hardening rates be- Fig. 3a shows the SEM microstructure of Fe–21Mn–3Al–3Si steel in
tween samples deformed at the temperatures 20 °C and 0 °C were rather the initial state. Nearly equiaxed austenite grains of various size, fre-
small. In both cases, the stress values achieved after 10% strain were quently exceeding even 100 μm, with numerous annealing twins are
within a relatively narrow range of 480–500 MPa. Only the reduction of visible. As may be seen from Fig. 3b, i.e. the inverse pole figure (IPF)
deformation temperatures to −50 °C and then to −150 °C caused a map, austenite grains showed almost random orientation distribution
significant increase in stress up to about 560 MPa and 620 MPa re- after the preliminary treatment. It should be noted however, based on
spectively. the EBSD map in Fig. 3c, that the applied heat treatment resulted in the
appearance of some amount of the α-ferrite within the microstructure
of the initial material. A selective quantitative assessment showed, that
3.2. X-ray analysis the average amount of ferrite (α) is approximately up to about 3% of
the volume fraction. The α-ferrite grains had much smaller size in
X-ray diffraction patterns registered for the examined steel in the comparison to austenite, reaching only occasionally up to about 30 μm
in diameter, and were rather non-uniformly distributed within the
Table 2 austenitic matrix. The crystallographic orientation relationship of
Mechanical properties of Fe-Mn-Al-Si steel at different deformation tempera- Kurdjumov-Sachs (K–S) was found to dominate between the austenite
tures. (γ) and ferrite (α) grains, {111}γ‖{101}α and < 101 > γ‖ < 111 > α,
Temperature of tensile test UTS Total elongation Microhardness, HV0.1
which in Fig. 3c is marked with a thin white line.
MPa % After 10% of tensile strain the structural effects of deformation
appeard in austenite grains for all the temperature variants (Figs. 4–7).
20 °C 680 57 244 These effects are in the form of thin deformation bands from at least one
0 °C 742 65 256
slip system. Taking into account the calculated SFE value of the steel
−50 °C 845 58 270
−150 °C 1017 54 294 (14.26 mJ/m2 - see section 3.4) this should be a slip of dissociated
partial dislocations. At the temperature 20 °C a slight proportion of the

3
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

Fig. 2. X-ray diffractograms for Fe-Mn-Si-Al steel samples in the initial state (0%) and after 10% of tensile strain at four deformation temperatures (a), the expanded
plot the (111) and (200) for austenite and the (0002) and (10‾ 1 1) for ε-martensite peaks (b).

ε-martensite with a hexagonal close-packed structure appeared within clearly indicate at the reduction of austenite stability together with the
the microstructure of examined steel (Fig. 4, Table 3). However the fall of deformation temperature, and are in agreement with a well-
average fraction of this phase was very small and almost inconsiderable known experimental fact, that lowering the temperature results in a
in some places. As the temperature of deformation was reduced, to 0 °C, decrease in the SFE value [4,17,27]. In general it should be noted, that
−50 °C and finally −150 °C, the stability of austenite gradually de- in the steel under examination the strain induced γ (fcc) → ε (hcp)
creased and the martensitic transformation took place to a much greater transformation was initiated at low strains, regardless of the applied
extent. Starting from the temperature 0 °C, the contribution of strain deformation temperature.
induced ε-phase has considerably increased after 10% of tensile strain, In the case of the α′-martensite, it was difficult to determine the
reaching in places about 10.5% of the volume fraction. exact mean values of its volume fractions, because of the selective
Afterwards, the amount of the ε-phase exceeded 16% at the tem- character of the measurements, which differed considerably depending
perature −50 °C, and then increased to about 20% of volume fraction at on the place of measurement. Within the range of small strains, diverse
the temperature −150 °C (Table 3 and Figs. 4c–7c). These results results were obtained due to the low contents of α′-martensite and first

Fig. 3. Microstructure of Fe-Mn-Al-Si steel in the


initial state: band contrast image (a), IPF map (b),
EBSD phase map (red colour - γ-austenite, blue
colour - α-ferrite), the K-S relationship is marked
with a white line (c). (For interpretation of the re-
ferences to colour in this figure legend, the reader is
referred to the Web version of this article.)

4
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

Fig. 4. Microstructure of Fe-Mn-Al-Si steel after


10% of deformation at temperature 20 °C: band
contrast image (a), IPF map (b), EBSD phase map
(red colour - γ-austenite, blue - α-ferrite, yellow - ε-
martensite), the K-S and S-N relationships are
marked with a white and black lines respectively
(c). (For interpretation of the references to colour in
this figure legend, the reader is referred to the Web
version of this article.)

of all its local occurrence, mainly at the intersections of ε-phase bands. explanations of relatively very low volume fraction of α′-martensite at
At this point it should be emphasized that, according to a number of this level of deformation. Additionally, the places were also found,
observations (e.g. Bracke et al. [29]), the α′-phase embryos do not where the α′-martensite was formed within a single set of parallel bands
expand beyond the relatively small intersection areas of the ε-phase of the ε-phase (Fig. 6c). In general, this observation is consistent with
bands, where they are usually formed. This may be one of the the results of Olson and Cohen [13,30,31], who showed that the

Fig. 5. Microstructure of Fe-Mn-Al-Si steel after


10% of deformation at temperature 0 °C: band
contrast image (a), IPF map (b), EBSD phase map
(red colour - γ-austenite, blue - α-ferrite, yellow - ε-
martensite), the S-N relationship is marked with a
black line (c). (For interpretation of the references
to colour in this figure legend, the reader is referred
to the Web version of this article.)

5
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

Fig. 6. Microstructure of Fe-Mn-Al-Si steel after 10% of deformation at temperature −50 °C: band contrast image (a), IPF map (b), EBSD phase map (red colour - γ-
austenite, blue - α′-martensite, yellow - ε-martensite), the K-S, S-N and B relationships are marked with a white, black and green line respectively (c), pole figures
from the area marked with a rectangle on Fig. 6c for FCC, HCP and BCC phase (d). (For interpretation of the references to colour in this figure legend, the reader is
referred to the Web version of this article.)

nucleation of the α′-martensite may proceed within the areas con- intersections, which lead to the formation of α′-martensite embryos and
taining bundles or packets of stacking faults, the ε-phase or deformation then small islands.
twins. However, an exact explanation of this result will be given later in Fig. 6c shows the formation of α′-martensite in austenite grains,
this chapter. where only a single system of the ε-bands was visible. According to
Another complication in determining the real amount of this phase Olson and Cohen [13,31], two shear displacements are required to form
resulted from the difficulty in distinguishing between the α′-martensite the α′-martensite. This principle was based on the model for the (fcc→
and α-ferrite. Because of the very low carbon contents (0.02 wt%C) α′- bcc) lattice transition, originally suggested by Bogers and Burgers [32].
martensite had almost the same structure as α-ferrite, i.e. the body It indicates that the shear necessary for this transition should be divided
centered cubic structure (bcc), instead of the body centered tetragonal between two invariant plain strains that can take place sequentially or
(bct) as in high-carbon steels. Thus, the only major differences between simultaneously. Hence, attempts to explain the formation of α′-mar-
both phases consisted in the morphology as well as nucleation sites and tensite in cases of adjacent and non-intersecting bands always use ad-
the resulting places of α′-martensite formation. ditional sources of stresses operating in non-parallel systems (e.g. Ryś
As far as the ε-martensite is concerned, its contents within the range and Cempura [33], Fujita and Katayama [14], etc.).
of small strains is much larger and even the local character of the The explanation proposed in this case indicates at the onset of a
measurements indicates at clear upward trend of its volume fraction secondary compatibility slip within the austenite grains. Therefore, in
with decreasing temperature. This is especially well visible in spite of the fact that only a single deformation system of the ε-bands
Figs. 5c–7c, i.e. on EBSD phase maps, where the increasing number of ε- was detected, there are sufficient stresses acting in still non-operating
martensite bands (yellow colour) appeared on the background of un- (not revealed) secondary systems capable to induce nucleation and then
stable austenite (red colour). The lower deformation temperature, the formation of the α′-martensite. It should be emphasized that the shapes
higher density of ε-bands, and the more frequent their mutual of the just formed α′-areas are extended parallel to the traces of planes

6
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

orientation relationship between ε-martensite and α′-martensite, which


obeys the Burgers relationship i.e. {0001}ε‖{110}α′ and < 1120 >
ε‖ < 111 > α′, is marked with a thin green line (see Fig. 6c and d and
7c,d).
Fig. 7b and c displays the EBSD and IPF maps together with the
enlarged fragments for samples deformed at −150 °C. It is well visible
that α′-martensite formed islands exactly at the intersecting regions of
the ε-bands. Similarly the enlarged fragment of the EBSD map for
sample deformed at −50 °C indicates, that α′-martensite areas were
located exactly inside the bands of ε-phase (Fig. 6c). From these ob-
servations unambiguously results, that in examined steel austenite was
transformed into α′-martensite according to the sequence γ (fcc) → ε
(hcp) → α' (bcc).

4. Summary

The chemical composition of the steel under examination is a ty-


pical example of currently applied compositions for high-manganese
steels (e.g. Fe–Mn–Al–Si, Fe–Mn–C, Fe–Mn–Al–C, etc.). It includes high
contents of manganese with smaller additions of silicon and aluminum,
and a low amount of carbon, namely; Mn (21.2 wt%), Si (< 3 wt%), Al
(< 3 wt%) and C (0.02 wt%). In general, these steels have low stacking
fault energy at ambient temperature with SFE typically within the range
up to about 40 mJ/m2 [35,36]. Due to the chemical composition their
microstructure usually consist of single-phase austenite, which should
be a stable phase at room temperature and during further plastic de-
formation. In the case of the investigated steel the calculated SFE value
is 14,26 mJ/m2 [36,37]. Nevertheless, the microstructure of examined
steel contained small amount (roughly up to 3%) of bcc α-ferrite in the
initial state, i.e. after the preliminary treatment. Similar initial micro-
structure, i.e. containing small amount of bcc-phase (about 7%),
showed the 20Mn-3Si-3 Al-0.045C steel studied by Shen et al. [38]. The
chemical composition of that steel was very similar to the steel under
examination, except the carbon content which was more than twice as
high. Carbon is a very effective austenite stabilizer, and this fact might
Fig. 7. Microstructure of Fe-Mn-Al-Si steel after 10% of deformation at tem-
explain the occurrence of both mechanisms during deformation of that
perature −150 °C: band contrast image (a), IPF map (b), EBSD phase map (red
steel, i.e. twinning and strain induced phase transformation γ (fcc) → ε
colour - γ-austenite, blue - α-ferrite and α′-martensite, yellow - ε-martensite),
and IPF map from the area marked with a rectangle on Fig. 7b, the K-S, S-N and (hcp) → α' (bcc). On the other hand, the results of X-ray diffraction
B relationships are marked with a white, black and green line respectively (c), analysis and microstructure observations indicate that during de-
pole figures from the area marked with a rectangle on Fig. 7c for FCC, HCP and formation the austenite in examined 21Mn-3Si-3Al-0.02C steel turned
BCC phase (d). (For interpretation of the references to colour in this figure out to be metastable and underwent a strain-induced phase transfor-
legend, the reader is referred to the Web version of this article.) mation into martensite. The interpretation of strain-induced martensitic
transformation is based on the two ways of austenite to martensite
Table 3 transition, which concern hcp ε-martensite and bcc (or bct) α′-mar-
Phase composition of Fe-Mn-Al-Si steel in the initial state and after tensile tensite. The observed austenite instability refers not only to deforma-
deformation of 10% at reduced temperatures, estimated based on several local tion carried out at reduced temperatures, since a certain amount of ε-
EBSD measurements. martensite appeared within the microstructure of the steel already after
10% of tensile strain at ambient temperature of 20 °C. With the re-
Deformation FCC (γ) HCP (ε) BCC (α/α′) Non indexed
temperature phase [%] phase [%] phase [%] points [%] duction of deformation temperature, a gradual increase in the volume
fraction of the ε-phase was observed (Figs. 5c–7c and Table 3), up to
Initial state 97 – 3.0 – about 20% at −150 °C. It should be emphasized, that in the range up to
20 °C 96 0.8 3.1 0.1
10% of tensile strain, the amount of the α′-martensite was at the same
0 °C 86.8 10.4 2.4 0.6
−50 °C 80.8 16.4 2.5 0.3 time rather small. Despite the fact, that although in the steel under
−150 °C 74.6 20.0 4.9 0.5 examination the γ (fcc) → ε (hcp) → α' (bcc) transformation sequence
was observed, a deformation behaviour of the steel within the range of
small strains was dominated by strain-induced γ (fcc) → ε (hcp) mar-
of potential secondary systems. tensitic transformation. The lower the deformation temperature, the
It is well known from the literature, that hexagonal martensite is larger the amount of austenite transformed into ε-martensite and in
preferentially formed in the {111} planes of fcc austenite [34]. It is consequence, the higher degree of strain hardening was found. There-
obvious that the arrangement of atoms in {111}γ and (0001)ε planes of fore, the observed difference in strain-hardening is supposed to be first
both structures as well as along the respective directions < 110 > γ of all the result of the amount of strain-induced ε-martensite. The close-
and < 1120 > ε is identical. packed hexagonal structure is highly anisotropic and the easy slip is
Thus, the orientation relationship found between γ-austenite and ε- essentially possible only on the basal planes (0001) in three slip di-
martensite obeyed the Shoji-Nishiyama (S–N) relationship rections < 11‾ 2 0 > . As a result the number of slip systems showing low
{111}γ‖{0001}ε and < 110 > γ‖ < 1120 > ε, which is marked with a values of critical resolved shear stresses (CRSS) is the least in the ε-
thin black line on enlarged fragments in Figs. 6c and 7c. In turn the martensite and the remaining potential slip planes are characterized by

7
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

much higher values of CRSS. That is why, the ε-phase is usually re- 5. Concluding remarks
ported as a strong barrier for further plastic flow.
As mentioned before, the formation mechanism of deformation During reduction of deformation temperature in tensile tested high-
bands in this type of alloys is associated with the accumulated motion manganese Fe–Mn–Si–Al steel, a marked increase in the ultimate tensile
of planar defects in the form of so called extended dislocations, i.e. strength was observed and simultaneously a relatively small decrease in
dissociated dislocations coupled by the SF ribbons. Depending on the total elongation. It should be emphasized that the steel showed a very
SFE value and the resulting manner of stacking fault overlapping (in good combination of strength and plastic properties for each of the
other words, SF spatial arrangement) the bands of ε-martensite or twin applied test temperatures from the range 20 °C to −150 °C.
bands may be formed, or simply dense stacking fault packets Due to the small SFE value of the steel, the dominant deformation
[13,15,29,39]. The deformation behavior of high-manganese steels may mechanism at the onset of yielding was slip of dissociated dislocations.
considerably change depending on the applied temperature range and Just afterwards, still within the range of small strains, the strain-in-
the exact chemical composition. It follows from the fact that both fac- duced phase transformation γ (fcc) → ε (hcp) was increasingly occur-
tors, i.e. the temperature and the chemical composition, determine the ring during lowering of the deformation temperature.
SFE value and in consequence the operating deformation mechanisms, The carried out analysis indicates that in the steel under examina-
which result in a specific microstructure and final mechanical proper- tion the deformation-induced phase transformations occurred at low
ties. temperatures in accordance with the following sequence γ (fcc) → ε
For example, Choi et al. [10] examined the microstructure of the (hcp) → α' (bcc). In the range of small strains a considerable increase in
austenitic steel Fe–24Mn-0.12C tensile tested at various temperatures work hardening with decreasing temperature corresponded to a sig-
within the range 7 °C–277 °C. In the case of samples deformed at tem- nificant increase in the amount of ε-martensite within the micro-
peratures ranging from 7 °C to 57 °C, both, the strain induced ε-mar- structure of the steel.
tensite as well as deformation twins were found within the micro- Lowering of the deformation temperature had a very significant
structure of the steel. On the other hand, in the case of tensile tests influence on increasing the volume fraction of ε-martensite with a
conducted at higher temperatures, i.e. between 127 °C and 227 °C, the hexagonal structure. This effect was a consequence of the decrease in
twinning mechanism prevailed resulting in domination of deformation the SFE value and the reduction of austenite stability at low tempera-
twins within the microstructure. In turn Huang et al. [19] investigated tures.
mechanical properties of an austenitic Fe–23Mn–2Si–2Al (wt.%) steel, After 10% of tensile strain the volume fractions of ε-phase increased
with small addition of niobium, as a function of deformation tem- up to about 16% and 20% at the temperatures reduced to −50 °C and
perature, within the range 20 °C to −75 °C. The microstructure ob- −150 °C respectively. On the other hand the volume fraction of α′-
servations revealed a large number of twins induced by deformation martensite was rather slight and it was frequently formed at the in-
above −60 °C. As a result the steel achieved a relatively high total tersections of the ε-phase bands from two non-parallel sets.
elongations of about 80% and the increase of the ultimate tensile Additionally, formation of α′-martensite was also observed at the in-
strength from 660 to 750 MPa, which was the effect of twinning-in- tersections of ε-bands with non-parallel secondary slip systems.
duced plasticity (TWIP). At temperatures below −60 °C, the stability of Preferential crystallographic orientation relations were detected,
the austenite decreased and martensitic transformations occurred, γ dominating between γ-austenite and the strain induced ε- and α′-
(fcc) → ε (hcp) → α' (bcc), as a result of TRIP effect. Due to transfor- phases, i.e.: Shoji-Nishiyama (S–N) - {111}γ‖{0001}ε
mation-induced plasticity, total elongations attained 87% and ultimate and < 110 > γ‖ < 1120 > ε and Kurdjumov-Sachs (K–S) -
tensile strength reached up to 810 MPa. {111}γ‖{110}α′ and < 110 > γ‖ < 111 > α′ respectively. In con-
During lowering of deformation temperature the rate of transfor- sequence, between ε-martensite and α′-martensite, the Burgers re-
mation in the steel under examination and the resulting volume fraction lationship i.e. {0001}ε‖{110})α′ and < 1120 > ε‖ < 111 > α′ was
of ε-martensite bands considerably increased. Therefore, it should be observed.
assumed that the SFE value of the examined steel significantly de-
creased, despite the relatively high Mn contents (about 21 wt%). The Acknowledgements
results obtained by Li et al. [40] indicate that suitable content of Mn
increased the stability of austenite as well as the ε-martensite, because This work was sponsored by the National Science Centre, Poland
the SFE value of austenite increases with Mn content. The authors found under Contract no 2011/01/D/ST8/03905.
that α′-martensitic transformation was rather difficult in steel having The authors wish to thank Professor Piotr Bała for offering facilities
~19 wt% of Mn because of the high stability of ε-martensite. in AGH-UST Academic Centre for Materials and Nanotechnology, which
Simultaneously, it was frequently observed, that the intersecting made possible to conduct a significant part of the presented research.
bands of ε-martensite lead to strong local distortion of the austenite
lattice. According to a number of authors, such a situation may be a References
source of driving force for the formation of the α′-martensite
[12,19,29,34,41]. Therefore the deformation behaviour in the course of [1] O. Grässel, L. Krüger, G. Frommeyer, L.W. Meyer, High Strength Fe-Mn-C-(Al-Si)
strain induced transformations, which are engaging the three phases TRIP/TWIP steels development-properties-application, Int. J. Plast. 16 (2000)
1391–1409.
(i.e. γ-austenite, ε- and α′-martensite), is frequently explained in the [2] W. Bleck, K. Phiu-on, C. Herring, G. Hirt, Hot workability of as-cast high manganese
following way. The γ (fcc)→ε (hcp) transformation is considered to high-carbon steels, Steel Research. Steel Res. Int.. 78 (7) (2007) 536–545.
increase the strength of deformed steel and, on the other hand, the [3] O. Bouaziz, S. Allain, C.P. Scott, P. Cugy, D. Barbier, High manganese austenitic
twinning induced plasticity steels: a review of the microstructure properties re-
occurrence of the ε (hcp) → α’ (bcc) transformation is supposed to lationships, Curr. Opin. Solid State Mater. Sci. 15 (2011) 141–168 https://doi.org/
decrease the local internal stresses resulting from the intersection of ε- 10.1016/j.cossms.2011.04.002.
bands (Li et al. [40]). It is assumed that intersections of ε-bands with [4] G. Frommeyer, U. Brüx, P. Neumann, Supra-ductile and high-strength manganese-
TRIP/TWIP steels for high energy absorption purposes, ISIJ Int. 43 (3) (2003)
non-parallel secondary slip systems, often observed in the examined 438–446.
steel at the initial stage of deformation, may also have the same effect. [5] J. Kowalska, W. Ratuszek, M. Witkowska, A. Zielińska-Lipiec, T. Tokarski,
In general, martensitic transformation taking place during deformation Microstructure and texture characteristics of the metastable Fe-21Mn-3Si-3Al alloy
after cold deformation, J. Alloy. Comp. 643 (2015) S39–S45 https://doi.org/10.
of the examined high manganese steel affected in a very significant way
1016/j.jallcom.2015.04.086.
the work hardening behaviour and mechanical properties through the [6] F.Y. Lu, P. Yang, L. Meng, F. Cui, H. Ding, Influences of thermal martensites and
TRIP effect, starting from the range of relatively small strains. grain orientations on strain-induced martensites in high manganese TRIP/TWIP
steels, J. Mater. Sci. Technol. 27 (3) (2011) 257–265.

8
J. Kowalska, et al. Materials Science & Engineering A 767 (2019) 138406

[7] R.E. Schramm, R.P. Reed, Stacking fault energies of seven commercial austenitic hot ductility of high alloy Fe-xMn-C-yAl austenite TWIP steels, Mater. Sci. Eng. A
stainless Steels, Metall. Trans. 6A (1975) 1345–1351. 708 (2017) 360–374 https://doi.org/10.1016/j.msea.2017.10.001.
[8] Y.S. Han, S.H. Hong, The effect of Al on mechanical properties and microstructures [26] A. Hamada, A. Khosravifard, D. Porter, L.P. Karjalainen, Physically based modeling
of Fe-32Mn-12Cr-xAl-0.4C cryogenic alloys, Mater. Sci. Eng. A 222 (1997) 76–83. and characterization of hot deformation behavior of twinning-induced plasticity
[9] M. Ghasri-Khouzani, J.R. McDermid, Effect of carbon content on the mechanical steels bearing vanadium and niobium, Mater. Sci. Eng. A 703 (2017) 85–96 https://
properties and microstructural evolution of Fe-22Mn-C steels, Mater. Sci. Eng. A doi.org/10.1016/j.msea.2017.07.038.
621 (2015) 118–127 https://doi.org/10.1016/j.msea.2014.10.042. [27] X-h. Fang, P. Yang, F-y. Lu, L. Meng, Dependence of deformation twinning on grain
[10] H.C. Choi, T.K. Ha, H.C. Shin, Y. Chang, The formation kinetics of deformation twin orientation and texture evolution of high manganese TWIP steels at different de-
and deformation induced ε-martensite in an austenitic Fe-C-Mn steel, Scr. Mater. 40 formation temperatures, J. Iron Steel Res. Int. 18 (11) (2011) 46–52.
(1999) 1171–1177. [28] M. Koyama, T. Sawaguchi, T. Lee, C.S. Lee, K. Tsuzaki, Work hardening associated
[11] L. Remy, A. Pineau, Twinning and strain-induced F.C.C.→H.C.P. transformation in with ε-martensitic transformation, deformation twinning and dynamic strain aging
the Fe-Mn-Cr-C system, Mater. Sci. Eng. 28 (1977) 99–107. in Fe-17Mn-0.6C and Fe-17Mn-0.8C TWIP steels, Mater. Sci. Eng. A 528 (2011)
[12] H. Ding, H. Ding, D. Song, Z. Tang, P. Yang, Strain hardening behavior of a TRIP/ 7310–7316 https://doi.org/10.1016/j.msea.2011.06.011.
TWIP steel with 18.8% Mn, Mater. Sci. Eng. A 528 (2011) 868–873 https://doi.org/ [29] I. Bracke, G. Mertens, J. Penning, B.C. De Cooman, M. Liebeherr, N. Akdut,
10.1016/j.msea.2010.10.040. Influence of phase transformations on the mechanical properties of high-strength
[13] G.B. Olson, M. Cohen, Kinetics of strain-induced martensitic nucleation, Metall. austenitic Fe-Mn-Cr steel, Metall. Mater. Trans. A 37A (2006) 307–317.
Mater. Trans. A 6A (1975) 791–795. [30] G.B. Olson, M. Cohen, A mechanism for the strain-induced nucleation of martensitic
[14] H. Fujita, T. Katayama, In-situ observation of strain-induced γ→ε→α’ and γ→α’ transformations, J. Less Common. Met. 28 (1972) 107–118.
martensitic transformations in Fe-Cr-Ni alloys, Mater. Trans. 33 (1992) 243–252. [31] G.B. Olson, M. Cohen, A general mechanism of martensitic nucleation: Part II. fcc→
[15] S. Kibey, J.B. Liu, D.D. Johnson, H. Sehitoglu, Predicting twinning stress in fcc bcc and other martensitic transformations, Metall. Trans. A 7A (1976) 1905–1914.
metals: linking twin-energy pathways to twin nucleation, Acta Mater. 55 (2007) [32] A.J. Bogers, W.G. Burgers, Partial dislocations on the {110} planes in the BCC
6843–6851 https://doi.org/10.1016/j.actamat.2007.08.042. lattice and the transition of the FCC into the BCC lattice, Acta Metall. 12 (1964)
[16] A. Dumay, J.P. Chateau, S. Allain, S. Migot, O. Bouaziz, Influence of addition ele- 255–261.
ments on the stacking-fault energy and mechanical properties of an austenitic Fe- [33] J. Ryś, G. Cempura, Microstructure and deformation behaviour of metastable du-
Mn-C steel, Mater. Sci. Eng. A 483–484 (2008) 184–187 https://doi.org/10.1016/j. plex stainless steel at high rolling reductions, Mater. Sci. Eng. A 700 (2017)
msea.2006.12.170. 656–666 https://doi.org/10.1016/j.msea.2017.06.022.
[17] A.S. Hamada, L.P. Karjalainen, M.C. Somani, The influence of aluminum on hot [34] W. Zhang, Z. Liu, Z. Zhang, G. Wang, The crystallographic mechanism for de-
deformation behavior and tensile properties of high-Mn TWIP steels, Mater. Sci. formation induced martensitic transformation observed by high resolution trans-
Eng. A 467 (2007) 114–124 https://doi.org/10.1016/j.msea.2007.02.074. mission electron microscope, Mater. Lett. 91 (2013) 158–160 https://doi.org/10.
[18] Y.F. Shen, N. Jia, L. Zuo, R.D.K. Misra, Softening behavior by excessive twinning 1016/j.matlet.2012.09.086.
and adiabatic heating at high strain rate in a Fe-20Mn-0.6C TWIP steel, Acta Mater. [35] R. Xiong, H. Peng, H. Si, W. Zhang, Y. Wen, Thermodynamic calculation of stacking
103 (2016) 229–242 https://doi.org/10.1016/j.actamat.2015.09.061. fault energy of the Fe–Mn–Si–C high manganese steels, Mater. Sci. Eng. A 598
[19] B.X. Huang, X.D. Wang, Y.H. Rong, L. Wanga, L. Jinc, Mechanical behavior and (2014) 376–386 https://doi.org/10.1016/j.msea.2014.01.046.
martensitic transformation of an Fe–Mn–Si–Al–Nb alloy, Mater. Sci. Eng. A 438–440 [36] S. Allain, J.-P. Chateau, O. Bouaziz, S. Migot, N. Guelton, Correlations between the
(2006) 306–311 https://doi.org/10.1016/j.msea.2006.02.150. calculated stacking fault energy and the plasticity mechanisms in Fe–Mn–C alloys,
[20] Z.C. Luo, M.X. Huang, Revisit the role of deformation twins on the work-hardening Mater. Sci. Eng. A 387–389 (2004) 158–162 https://doi.org/10.1016/j.msea.2004.
behavior of twinning-induced plasticity steels, Scr. Mater. 142 (2018) 28–31 01.059.
https://doi.org/10.1016/j.scriptamat.2017.08.017. [37] O. Grässel, G. Frommeyer, C. Derder, H. Hofmann, Phase transformations and
[21] S. Allain, J.-P. Chateau, O. Bouaziz, A physical model of the twinning-induced mechanical properties of Fe-Mn-Si-A1 TRIP-steels, J. Phys. IV France 7 (1997) C5-
plasticity effect in a high manganese austenitic steel, Mater. Sci. Eng. A 387–389 383-C5388 https://doi.org/10.1051/jp4:1997560.
(2004) 143–147 https://doi.org/10.1016/j.msea.2004.01.060. [38] Y.F. Shen, Y.D. Wang, X.P. Lui, X. Sun, R. Lin Peng, S.Y. Zhang, L. Zuo, P.K. Liaw,
[22] Y.Z. Li, Z.C. Luo, Z.Y. Liang, M.X. Huang, Effect of carbon on strain-rate and tem- Deformation mechanisms of a 20Mn TWIP steel investigated by in situ diffraction
perature sensitivity of twinning-induced plasticity steels: modeling and experi- and TEM, Acta Mater. 61 (2013) 6093–6106 https://doi.org/10.1016/j.actamat.
ments, Acta Mater. 165 (2019) 278–293 https://doi.org/10.1016/j.actamat.2018. 2013.06.051.
11.048. [39] H. Fujita, S. Ueda, Stacking faults and F.C.C. (γ)→H.C.P. (ε) transformation in 18/8-
[23] Z.Y. Liang, Z.C. Luo, M.X. Huang, Temperature dependence of strengthening me- type stainless steel, Acta Metall. 20 (1972) 759–767.
chanisms in a twinning-induced plasticity steel, Int. J. Plast. 116 (2019) 192–202 [40] X. Li, L. Chen, Y. Zhao, R.D.K. Misra, Influence of manganese content on ε-/α′-
https://doi.org/10.1016/j.ijplas.2019.01.003. martensitic transformation and tensile properties of low-C high-Mn TRIP steels,
[24] F.X. Yin, H. Xia, J.H. Feng, M.H. Cai, X. Zhang, G.K. Wang, T. Sawaguchi, Mater. Des. 142 (2018) 190–202 https://doi.org/10.1016/j.matdes.2018.01.026.
Mechanical properties of an Fe-30Mn-4Si-2Al alloy after rolling at different tem- [41] R.T. van Tol, J.K. Kim, L. Zhao, J. Sietsma, B.C. De Cooman, α’-martensite forma-
peratures ranging from 298 to 1073 K, Mat er, Sci. Eng. A 725 (2018) 127–137 tion in deep-drawn Mn-based TWIP steel, J. Mater. Sci. 47 (2012) 4845–4850
https://doi.org/10.1016/j.msea.2018.03.079. https://doi.org/10.1007/s10853-012-6345-y.
[25] H. Liu, J. Liu, B. Wu, Y. Shen, Y. He, H. Ding, X. Su, Effect of Mn and Al contents on

You might also like