Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Science and Engineering A 496 (2008) 417–424

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

Microband-induced plasticity in a high Mn–Al–C light steel


Je Doo Yoo, Kyung-Tae Park ∗
Division of Advanced Materials Science & Engineering, Hanbat National University, Taejon 305-719, South Korea

a r t i c l e i n f o a b s t r a c t

Article history: Room temperature tensile behavior of a high Mn–Al–C steel in the solid solution state was correlated to
Received 16 April 2008 the microstructures developed during plastic deformation in order to clarify the dominant deformation
Received in revised form 26 May 2008 mechanisms. The steel was fully austenitic with a fairly high stacking fault energy of ∼85 mJ/m2 . The
Accepted 26 May 2008
tensile behavior of the steel was manifested by an excellent combination of strength and ductility over
80,000 MPa% in association with continuous strain hardening to the high strain. In addition, the austenite
Keywords:
phase was very stable during deformation. The high stacking fault energy and firm stability of austenite
Steel
were attributed to the high Al content. In spite of the high stacking fault energy, deformed microstructures
Austenite
Deformation
exhibited the planar glide characteristics, seemingly due to the glide plane softening effect. In the process
Stacking fault energy of straining, the formation of crystallographic microbands and their intersections dominantly occurred.
Planar glide Microbands consisting of geometrically necessary dislocations led to the high total dislocation density
Microbands state during deformation, resulting in continuous strain hardening. This microband-induced plasticity is
to be the origin of the enhanced mechanical properties of the steel.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction ing [4–6]. However, weight savings of TWIP steels are not effective
because of the atomic weight of Mn, the major alloying element of
Development of super high strength steels with good ductil- TWIP steels, comparable to that of Fe and low alloying of the light
ity and toughness has long been pursued in steel industries. As elements such as Al.
a result, especially for automotive sheet steels, several advanced Several recent investigations [4,7,8] revealed that the Al addition
steels have been introduced such as interstitial free steels, dual up to ∼10 wt.% into the high Mn austenitic steels is beneficial for
phase steels, transformation-induced plasticity aided steels, etc. achieving not only remarkable weight savings but also mechanical
Although their strength and corresponding ductility vary in a properties comparable to or better than those of TWIP steels. These
wide range depending on the constituent phases and chemical effects mainly result from the SFE increase with increasing the Al
composition, a combination of strength and ductility (i.e. tensile content in the high Mn austenitic steels and the corresponding
strength (TS) × elongation) of commercially available automotive change of the deformation mode from TWIP to dislocation slip [4].
sheet steels mostly coalesces into 15,000–20,000 MPa% in the plot It is also reported that the C addition into the high Mn–Al austenitic
of tensile strength against elongation [1]. steels provides further strengthening by ␬ -carbides (ordered L 12
In response to the recent skyrocketing fuel cost and more strict structure of (Fe,Mn)3 AlCx ) precipitation [7–9].
regulations for safety and CO2 emission, automotive industries At present, little attention has been paid to the microstructural
demand more advanced steels over 40,000 MPa% with the light evolution during plastic deformation of high Mn austenitic steels
specific weight. At present, twin-induced plasticity (TWIP) aided with the increased SFE by adding the high Al content and therefore
steels satisfy such mechanical demand by the tensile strength of the origin of the attractive combination of strength and ductility
∼900 MPa and extended ductility over ∼50% [2,3]. The extraordi- of these steels is unclear. In the present investigation, the corre-
nary mechanical properties of TWIP steels are attributed to the fact lation between room temperature tensile behavior and deformed
that twins are dynamically formed during deformation of austen- microstructure of a high Mn–Al–C steel was analyzed in order to
ite with the high Mn content due to a relatively low stacking fault understand the plastic deformation mechanisms of the steel, espe-
energy (SFE) of 20–40 mJ/m2 , and these twins act as the effective cially focusing on its strain hardening behavior.
obstacles for dislocation movement leading to high strain harden-
2. Experimental

∗ Corresponding author. Tel.: +82 42 821 1243; fax: +82 42 821 1592. Twenty-five-kilogram ingots of a high Mn–Al–C steel were
E-mail address: ktpark@hanbat.ac.kr (K.-T. Park). prepared by induction melting in an argon atmosphere. After

0921-5093/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2008.05.042
418 J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424

considering the annealing twin boundaries. The stability of austen-


ite of the steel during preparation and deformation was examined
by X-ray diffraction (XRD, Rigaku, D/MAX 2500H). Submicrostruc-
tures of deformed specimens were observed by transmission
electron microscopy (TEM, JEOL 1010) operating at 200 kV. Thin
foils for TEM observation were prepared by a twin-jet polishing
technique using a mixture of 10% perchloric acid and 90% methanol
at −35 ◦ C with an applied potential of 20 V.

3. Results

3.1. Initial microstructure and stacking fault energy

The band contrast image of the microstructure of the steel after


solution treatment followed by quenching is shown in Fig. 1a. The
annealing twins prevailed in most grains. The grain size was 15
and 37 ␮m with/without considering the annealing twin bound-
aries, respectively. The XRD analysis (Fig. 1b) and the phase map
by EBSD analysis (not shown here) revealed that the steel was fully
austenitic.
Al is known to be the most effective on increasing the SFE of the
high Mn austenitic steels. Considering the SFE ( ) as the Gibbs free
energy (G␥→␧ ) required for the formation of ␧ martensite plate in
the austenite (␥) matrix as suggested by Hirth [10], the SFE can be
estimated by

 = 2G␥→␧ + 2 (1a)

Fig. 1. Microstructural characteristics of Fe–28Mn–9Al–0.8C steel after solution


treatment at 1000 ◦ C for 1 h followed by quenching: (a) Fe-SEM band contrast image
and (b) the XRD profile.

homogenization at 1200 ◦ C for 2 h, the ingot was hot-rolled to the


plate of 12 mm thickness: hot rolling finish temperature was 900 ◦ C.
The hot-rolled plate was annealed at 1000 ◦ C for 1 h and then cold-
rolled with a thickness reduction of 65%. The cold-rolled plate was
solution treated at 1000 ◦ C for 1 h and water-quenched. The chem-
ical composition of the steel after all above treatments was 0.79
C, 27.8 Mn, 9.1 Al, 0.003 P, 0.0037 S and the balance Fe in wt.%.
The density of the steel was measured as 6.78 g/cm3 , an apparent
reduction of ∼14% compared to pure Fe, by using a densitometry
(METTLER TOLEDO AT201) based on the Archimedes Principle.
Tensile specimens of a gage section of 25.4 mm × 2 mm × 6 mm
were machined from the water-quenched plate. Most room temper-
ature tensile tests were performed up to failure using a universal
testing machine (INSTRON model 4484) operating at constant
crosshead displacement rate with the initial strain rate of 10−3 s−1 .
Some tests were interrupted at predetermined strains to observe
the microstructures developed at the different strain levels.
Microstructures were optically examined by etching the
mechanically polished samples with 2% Nital. Since microstructure
of the high Mn austenitic steels usually contains a considerable
amount of annealing twins, an electron back scattered diffrac-
Fig. 2. (a) The representative engineering stress–strain curve of Fe–28Mn–9Al–0.8C
tion (EBSD) imaging system (Oxford INCA system) equipped in a
steel tested at room temperature with the initial strain rate of 10−3 s−1 . (b) The true
field emission gun scanning electron microscope (FEG-SEM, JSM stress–strain curve of (a) and the corresponding strain hardening rate as a function
6500F) was employed for the grain size measurement with/without of true strain.
J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424 419

Fig. 3. Optical micrographs showing the microstructures of Fe–28Mn–9Al–0.8C steel developed at different strains during room temperature tensile deformation: (a) 5%
(500×), (b) 5% (200×), (c) 40% (500×), (d) 40% (200×), (e) 60% (500×), (d) 60% (200×) (the stress axis is horizontal).

 4  1 
Tneel is the Neel temperature, ˇ is the magnetic moment depend-
= √ (1b)
3 a2 N ing on the composition, and RT has the usual meaning. From Eqs.
(2a)–2(c), G␥→␧ for ␥/␧ transformation of the binary austenitic
where  is the molar planar density of {1 1 1} planes,  is the ␥/␧
steel can be expressed by
interfacial energy, a is the lattice constant of ␥, and N is the Avo-
gadro number. G→ε can be obtained by applying the regular and
␥→␧ ␥→␧ ␥→␧ ␥→␧
subregular solution model [4,11]. For a single phase (␾) binary solu- G␥→␧ = Xi Gi + Xj Gj + Xi Xj ˝ij + Gmg (3)
tion consisting of the elements i and j, the Gibbs free energy (G ) is
given by By extending Eq. (3) for the multicomponent systems as proposed
␾ ␾ ␾ ␾ by Dumay et al. [8], G→ε of the present Fe–Mn–Al–C steel is given
G␾ = Xi Gi + Xj Gj + RT(Xi ln Xi + Xj ln Xj ) + Xi Xj ˝ij + Gmg (2a)
by
␾ ␾ ␾
ij
= L0 + L1 (Xi − Xj ) (2b)
␥→␧ ␥→␧ ␥→␧ ␥→␧
 T  GFe–Mn–Al–C = XFe GFe + XMn GMn + XAl GAl

Gmg =f RT ln(ˇ + 1) (2c) ␥→␧ ␥→␧ ␥→␧
Tneel +XC GC + XFe XMn ˝FeMn + XFe XAl ˝FeAl
where X is the atomic molar fraction, ˝ is the excessive mixing ␥→␧
+XFe XC ˝FeC + XMn XAl ˝MnAl
␥→␧
energy, Gmg is the Gibbs energy for the magnetic state of the phase
␥→␧ ␥→␧ ␥→␧
␾, L0 is a parameter depending on the temperature, L1 is a constant, +XMn XC ˝MnC + XAl XC ˝AlC + Gmg (4)
420 J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424

Table 1
The values used for the estimation of the SFE of a Fe–28Mn–9Al–0.8C steel by Eqs. (1) and (4) [4,8,12]

Parameters Values and functions

 2.94 × 10−5 (mol/m2 )


 9 (mJ/m2 )
␥→␧
GFe −2243.38 + 4.309T (J/mol)
␥→␧
GMn −1000 + 1.123T (J/mol)
␥→␧
GAl 2800 + 5T (J/mol)
␥→␧
GC −22,166 (J/mol)
␥→␧
˝FeMn 2180 + 532 (XFe − XMn ) (J/mol)
␥→␧
˝FeAl 3339 (J/mol)
␥→␧
˝FeC 42,500 (J/mol)
␥→␧
˝MnC 26,910 (J/mol)
ˇ 0.7XFe + 0.62XMn − 0.64XFe XMn − 4XC
ˇ␧ 0.62XMn − 4XC
T 298 (K)

TNeel 250 ln(XMn ) − 4750XMn XC − 6.2XAl + 720 (K)

TNeel 580XMn (K)
 
−1
 {(79 /140p) + (474/497)((1/p) − 1)(( 3
/6) + ( 6
/135) + ( 15
/600))} T
[5pt] f (T/TNeel ) 1− when = <
D 
TNeel
 
−5 −15 −25
{( /10) + ( /315) + ( /1500)} T
1− when = ) > 1, where p = 0.28 and D = 2.34 for FCC
D 
TNeel

From Eqs. (1a) and (4), the SFE and G→ε of the present 40%, all grains exhibited the intensive uni- or bi-directional strain
Fe–28Mn–9Al–0.8C steel at room temperature were estimated as markings (Fig. 3c) and grain elongation along the stress axis became
∼85 mJ/m2 and ∼1130 J/mol, respectively. The values and functions obvious (Fig. 3d). At a strain of 60%, the wavy strain markings
used for the estimation are listed in Table 1 [4,8,12]. appeared in some large grains (for instance, marked by ‘x’ in Fig. 3e)
and grains became more elongated (Fig. 3f). The strain markings on
3.2. Room temperature tensile behavior both sides of annealing twin boundaries yielded the exact mirror
image (‘x’ and ‘y’ in Fig. 3e), indicating that annealing twin bound-
3.2.1. Stress–strain curves aries acted as grain boundaries and the same slip system operated
Fig. 2a shows the representative engineering stress–strain curve on both sides. Up to failure, the entire strain markings ended at
of the steel. The steel exhibited continuous yielding and exten- either austenite grain boundaries or annealing twin boundaries.
sive strain hardening. The average yield strength (YS) and tensile The penetration of the strain markings across grain/twin bound-
strength from the three samples were ∼440 and ∼840 MPa, respec- aries was barely observed. Neither sample scale shear bands nor
tively. Exceptionally high ductility (ef ) of 100% was obtained. It those crossing several grains were found. As shown in Fig. 4, the
is usual to describe the combination of strength and correspond- fracture surface consisted of well-developed dimples, i.e. ductile
ing formability of automotive sheet steels in terms of the product failure rather than shear band induced brittle failure. No other
of TS and ef (TS × ef ) since they are mutually contradictory. The phases but austenite were found in the XRD analysis on the gage
typical values of (TS × ef ) of commercial low carbon automotive section of the failed samples (Fig. 4b), indicating the firm stability
sheet steels, TRIP steels and TWIP steels are ∼15,000, ∼20,000, of austenite of the present steel associated with the high G␥→␧ of
and ∼50,000 MPa%, respectively [1,2]. The value of (TS × ef ) of the ∼1130 J/mol.
present steel reached 84,000 MPa%, much higher than that of TWIP
steels. The true stress ()–strain (ε) curve of the steel is presented 3.2.2.2. TEM microstructures. Dislocation configuration of the sam-
in Fig. 2b with the corresponding strain hardening rate (d/dε). ple strained to 5% is shown in Fig. 5a. At this low strain level,
Strain hardening was not linear but its rate continuously increased substructural developments are manifested by dislocation pile ups
from the onset of plastic deformation to ε = 0.48, and then decreased on a single slip plane and slip along the principal crystallographic
yielding the Considere plastic instability at ε = 0.62 (equivalent to directions, which are the typical planar glide configurations [13]. At
the uniform elongation of 86%). The post-necking strain was ∼14%. a strain of 10% (Fig. 5b), no distinct cell formation was observed but a
Taylor lattice-like structure, a kind of low energy dislocation struc-
3.2.2. Deformed microstructures ture [14,15], was formed. The microstructure of the sample strained
3.2.2.1. Optical microstructures. Fig. 3 shows the optical micro- to a medium strain of 40% exhibited the well-developed microband
graphs of the steel deformed to the different strain levels. Even at a with distinct boundaries (marked ‘A’ in Fig. 5c). Microband in the
low strain of 5%, the parallel straight strain markings along the two forming process was also found (marked ‘B’ in Fig. 5c), exhibiting a
principal directions were observed in some grains, running through very high dislocation density between the ill-defined boundaries.
a whole grain (Fig. 3a). These strain markings were not appar- The microbands developed in no cell-forming FCC alloys such as
ently revealed in some grains and the grain shape remained nearly Al–5% Mg and 7030 Al are known to be oriented along the trace of
equiaxed at this low strain (Fig. 3b), implying the strong depen- a {1 1 1} slip plane [16,17]. In addition, slip became finer and more
dence of early stage yielding on the grain orientation. At a strain of intensive, and its bandwidth was reduced to less than 50 nm (the
J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424 421

purpose of comparison (Section 4.1) before discussing the defor-


mation characteristics of the present steel in detail (Section
4.2).

4.1. Stacking fault energy and deformation modes of austenitic


steels

When the SFE is less than the ␥/␧ interfacial energy in Eq.
(1a), G␥→␧ becomes negative. In such a case, the formation
of HCP ␧ martensite is energetically favorable by overlapping
of the intrinsic stacking faults on every second {1 1 1} plane
at the early stage of deformation [18,19]. By further straining,
the population of shear bands – here, ‘shear band’ is the col-
lective term for the planar defects formed by overlapping of
stacking faults [20] – increases and ␣ martensite is formed
at their intersections. The typical value of the ␥/␧ interfa-
cial energy is 10–20 mJ/mol [4–6,8]. Accordingly, the deformed
microstructure of austenitic steels having the SFE less than
∼20 mJ/mol, for instance 301 LN or 316 austenitic stainless steels
[20–22], exhibits the increasing ␣ martensite volume fraction with
increasing strain, resulting in the transformation-induced plastic-
ity.
By contrast, when the SFE is higher than the ␥/␧ interfacial
energy, G␥→␧ becomes positive and the formation of ␧ marten-
site is suppressed [18,19]. Instead, mechanical twinning is favorable
by displacement of partial dislocation on successive {1 1 1} planes
maintaining FCC structure. As mentioned earlier, these mechanical
twins enhance the strain hardenability, i.e. TWIP. Recently, by bal-
ancing all the forces acting on partial dislocations under the applied
stress, Byun [21] derived the critical stress ( T ) for occurrence of
twinning in polycrystalline materials such that


T = 6.14 (5)
b

Fig. 4. (a) SEM micrograph showing the dimpled ductile fracture surface of where b is the magnitude of the partial dislocation. At the
Fe–28Mn–9Al–0.8C steel failed at 100%. (b) The XRD profile taken from the gage stresses larger than  T , the stacking faults diverge and become
section of Fe–28Mn–9Al–0.8C steel failed at 100%.
unstable and therefore mechanical twinning occurs concurrently.
Contrarily, at the stresses below  T , the stacking faults have
inset of Fig. 5c). The submicrostructure of the sample failed at 100% the finite width depending on the applied stress and SFE, and
was dominated by the intersections of well-developed microbands deformation is achieved by partial and/or perfect dislocation glid-
(Fig. 5d). At some microband intersections, both sharp offset and ing.
smooth offset were found (indicated by arrows in Fig. 5d). Even at Meanwhile, Misra et al. [23] investigated the tensile-deformed
high strains up to failure, no cell formation occurred. microstructures of ultrafine/nano grained 301 LN austenitic stain-
less steel with SFE of ∼15 mJ/mol. They found that strain induced
4. Discussion ␣ martensite transformation associated with shear bands was
dominant in submicron grains of 500–1000 nm, and mechani-
The room temperature tensile characteristics of the present cal twinning more prevailed in ultrafine/nano grains less than
high Mn–Al–C steel are manifested by an excellent combination 500 nm. They rationalized the deformation mode change with
of high strength and extended ductility due to continuous strain the grain size such that, with decreasing the grain size, the acti-
hardening. The deformed microstructures exhibited the typical vation of the Frank-Read sources for shear band formation by
planar glide characteristics such as no cell formation, disloca- dislocation slip becomes more difficult, and hence deformation
tion pile ups on a single slip plane and formation of a Taylor is facilitated by twinning. They also mentioned the effect of SFE
lattice structure at low strains, crystallographic microband for- increment with decreasing the grain size on this deformation mode
mation at medium strains, and microband intersections at high change.
strains. More noticeably, neither ␣ martensite nor mechanical The SFE of the present steel is much higher than the ␥/␧
twins, which are often observed in the deformed microstruc- interfacial energy. In addition, its  T in Eq. (5) was estimated
ture of austenitic steels such as austenitic stainless steels and as higher as ∼3550 MPa by using  = 85 mJ/m2 and b = 0.147 nm
TWIP steels, was observed in the present steel. This is pri- [5]. Accordingly it is obvious that the formation of strain induced
marily attributed to the effects of the SFE on the deformation ␣ martensite and mechanical twins hardly occurs during plas-
mode: the SFE of the present steel ( = 85 mJ/m2 ) is much tic deformation of the present coarse grained (37 ␮m) austenitic
higher than that of austenitic stainless steels ( < 20 mJ/m2 ) steel at room temperature. Instead, plastic deformation of the
and TWIP steels (20 mJ/m2 <  < 40 mJ/m2 ). Accordingly, it is present steel is sure to be achieved by gliding of perfect dislo-
instructive to describe the effects of SFE on the deformation cations and/or partial dislocations with very narrow spacing as
mode of austenitic steels with the relatively low SFE for the observed.
422 J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424

Fig. 5. TEM micrographs showing the microstructures of Fe–28Mn–9Al–0.8C steel developed at different strains during room temperature tensile deformation: (a) 5%, (b)
10%, (c) 40% and (d) after failure at 100%.

4.2. Deformation characteristics of fully austenitic ing (SRO) in FCC solid solution alloys is mainly responsible for
Fe–28Mn–9Al–0.8C steel planar glide regardless of the value of SFE [13,25–27]: for exam-
ple, Table 1 in Ref. [13]. The leading dislocation moving through
4.2.1. Planar glide and glide plane softening lattice faces the energy barrier by SRO. The following dislocations
In general, for FCC materials, planar glide is explained to occur activated on the same single slip plane are to be piled up, for exam-
if they have a low SFE and so their extended partial dislocations ple Fig. 5a, and they help the leading dislocation overcome the SRO
are widely separated. In such a case, dislocations have the low energy barrier by destroying SRO itself. Since the order is in the
three-dimensional mobility and therefore insufficient Burgers vec- short range, the destroyed order is hard to be restored. Then the
tor directions for cell formation, tending to planar glide. By contrast, following dislocations glide more easily over the destroyed SRO
if a SFE is high enough, cross-slip of the screw component of region on that plane following the leading dislocation, i.e. glide
extended partials is easy to occur, tending to wavy slip forming a cell plane softening. Accordingly, plastic deformation is localized on
structure [24]. But, in spite of its high SFE, the present steel exhib- a single slip plane and exhibits the planar glide manner. Planar
ited the planar glide characteristics and no cell formation occurred glide tied to glide plane softening is reported to occur more pref-
even up to failure. This fact indicates that the SFE is not a critical erentially in the concentrated solid solutions [13,27]. The present
material parameter determining the slip mode, i.e. planar glide or Fe–28Mn–9Al–0.8C steel in the solid solution state is a concen-
wavy slip. Therefore, another origin of planar glide during plastic trated solid solution in which the total atomic mole fraction of the
deformation of the present steel is worth exploring. alloying elements reaches ∼0.45. Considering its relatively high
The several investigations previously reported that ‘glide plane SFE and concentrated alloying, it is not erroneous to conclude
softening’ in association with the occurrence of short range order- that planar glide during plastic deformation of the present steel
J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424 423

is attributed to the glide plane softening effect rather than the SFE ended at grain boundaries or annealing twin boundaries, and
effect. deformed microstructures clearly demonstrated the development
of microbands with strain. As discussed in Section 4.2, microbands
4.2.2. Microstructural evolution of planar glide materials during and their precursors, i.e. domain boundaries, are the structure con-
plastic deformation sisting of geometrically necessary dislocations. Their progressive
In the wavy slip materials, the imposed strain is mainly relaxed development with strain increases the total dislocation density and
by means of dislocation cell formation at low and medium strains therefore continuous strain hardening occurs [29]. Accordingly,
and non-crystallographic microbands are often formed at high the excellent combination of strength and ductility of the present
strains. By contrast, in the planar glide materials, a different strain high Mn–Al–C steel is primarily attributed to microband-induced
relaxation mechanism operates. At low strains, dislocations with a plasticity rather than shear band induced plasticity suggested by
lack of three-dimensional mobility rearrange themselves and form Frommeyer and Brüx [7]. In view of the engineering applications,
a finely organized structure called a Taylor lattice, for example by utilizing microband-induced plasticity and concentrated alloy-
Fig. 5b. A Taylor lattice is a homogeneous distribution of disloca- ing of a light element Al, the austenitic steels with the remarkably
tions with planar order about the glide planes such that the order reduced specific weight and enhanced mechanical properties can
of dislocations about the glide planes has an alternating sense be achieved as a strong candidate for advanced automotive sheet
[15,28]. In such a case, the stored energy is lowered by mutual steels.
screening due to their alternating sense. With increasing strain,
the crystallographic boundaries between the Taylor lattice domains 5. Conclusions
are formed—called domain boundary [15]. The domain boundary
is a single dislocation wall consisting of geometrically necessary 1. Fe–28Mn–9Al–0.8C steel in the solid solution state was fully
dislocations to accommodate the misfit orientation between the austenitic with abundant annealing twins. The stacking fault
Taylor lattice domains. By further increasing strain, the second energy of the steel was estimated as ∼85 mJ/m2 and its austenite
dislocation wall parallel to the first formed domain boundary stability was firm with G␥→␧ of ∼1130 J/mol.
appeared, for example Fig. 5c. The formation of the second dis- 2. The present fully austenitic steel exhibited an excellent combina-
location wall is energetically favored to screen the energy of the tion of strength and ductility over 80,000 MPa% with fairly high
first formed domain boundary. Accordingly, the second disloca- post-necking elongation of 14% at room temperature. The strain
tion wall also consists of geometrically necessary dislocations. The hardening rate of the steel during tensile deformation contin-
region between the first and second formed walls is usually called uously increased with increasing strain before yielding plastic
microband [14]. Since the low energy sites in no cell-forming mate- instability.
rials are grain boundaries, domain boundaries, microbands, etc., 3. The deformed microstructures of the steel exhibited the typi-
both domain boundaries and microbands end at these sites [29–31]. cal planar glide characteristics such as dislocation pile ups on
The length of some of them is reported to extend 50–100% of a single slip plane and formation of a Taylor lattice structure at
the grain diameter [16]. At high strains, the rotation of the Tay- low strains, crystallographic microband formation at medium
lor lattice domains increases. To accommodate this rotation, more strains, and microband intersections at high strains.
domain boundaries and microbands are formed and their spacing 4. Continuous strain hardening causing an excellent combination
decreases. Finally, their intersections cause grain subdivision, for of strength and ductility of the steel is to be originated by plas-
example Fig. 5d. The details of the microstructural evolution during ticity induced by microbands which represent a geometrically
plastic deformation of planar glide materials are well documented necessary structure with high dislocation density.
in Ref. [16].

4.2.3. Microband-induced plasticity Acknowledgment


Frommeyer and Brüx [7] investigated the mechani-
cal behavior of a variety of high Mn–Al–C steels ranging This study was supported by a grant from the Fundamental R&D
18/28Mn–9/12Al–0.7/1.2C. In case of Fe–28Mn–12Al–1.0C steel Program for Core Technology of Materials funded by the Ministry
close to the upper bound of alloying in their steels, they obtained of Commerce, Industry and Energy, Republic of Korea.
YS = 730 MPa, UTS = 1000 MPa and ef = 55% at room tempera-
ture. The strengths are higher than those of the present steel References
(YS = 440 MPa, UTS = 840 MPa), mainly due to the higher alloying [1] Ultra Light Steel Auto Body-Advanced Vehicle Technology (ULSAB-AVC)
contents and presence of ␬-carbides in their steel. The total overview report, IISI, 2002, p. 66, http://www.worldautoteel.org.
elongation is much smaller than that of the present one (∼100%). [2] O. Grässel, L. Krüge, G. Frommeyer, L.W. Meyer, Int. J. Plasticity 16 (2000)
1391–1409.
The lower elongation of their steel seems to be attributed to not [3] G. Frommeyer, U. Brüx, P. Neumann, ISIJ Int. 43 (2003) 438–446.
only higher strength but also the presence of BCC ferrite in the [4] S. Allain, J.-P. Chateau, O. Bouaziz, S. Miget, N. Guelton, Mater. Sci. Eng. A
austenite matrix, unlike the present steel with full austenite. 387–389 (2004) 158–162.
[5] S. Allain, J.-P. Chateau, O. Bouaziz, Mater. Sci. Eng. A 387–389 (2004) 143–147.
Frommeyer and Brüx [7] explained the mechanical behavior [6] S. Allain, J.-P. Chateau, D. Dahmonn, O. Bouaziz, Mater. Sci. Eng. A 387–389
of their high Mn–Al–C steels in term of shear band induced (2004) 272–276.
plasticity based on the limited TEM observation without pro- [7] G. Frommeyer, U. Brüx, Steel Res. Int. 77 (2006) 627–633.
[8] A. Dumay, J.-P. Chateau, S. Allain, S. Miget, O. Bouaziz, Mater. Sci. Eng. A 483–484
viding the microstructural evolution stages with increasing
(2008) 184–187.
strain. [9] W.K. Choo, J.H. Kim, J.C. Yoon, Acta Mater. 45 (1997) 4877–4885.
As aforementioned in Section 3.2.2.1, the present steel exhib- [10] J.P. Hirth, Metall. Trans. A 1 (1970) 2367–2374.
[11] Y.K. Lee, C.S. Choi, Metall. Mater. Trans. A 31A (2000) 355–360.
ited neither sample scale shear bands nor those crossing several
[12] L. Li, T.Y. Hsu, Calphad 21 (1997) 443–448.
grains—here, ‘shear band’ is the term describing the macroscopic [13] V. Gerold, H.-P. Karnthaler, Acta Metall. 37 (1989) 2177–2183.
narrow localized deformed zone compared to other regions of [14] D. Khulmann-Wilsdorf, Mater. Sci. Eng. A 113 (1989) 1–41.
the sample. In addition, it did not show the shear band induced [15] B. Bay, N. Hansen, D. Khulmann-Wilsdorf, Mater. Sci. Eng. A 113 (1989) 385–
397.
brittle failure but dimpled ductile failure after a prolonged post- [16] D.A. Hughes, Acta Metall. Mater. 41 (1993) 1421–1430.
necking strain. As shown in Figs. 3 and 5, the strain markings [17] A. Fjeldly, A. Sørey, H.J. Roven, Mater. Sci. Eng. A 300 (2001) 165–170.
424 J.D. Yoo, K.-T. Park / Materials Science and Engineering A 496 (2008) 417–424

[18] G.B. Olson, M. Cohen, Metall. Trans. A 7A (1976) 1897–1904. [26] T. Steffens, C. Schwink, A. Korner, H.-P. Karnthaler, Philos. Mag. A 56 (1987)
[19] G.B. Olson, M. Cohen, Metall. Trans. A 7A (1976) 1905–1914. 161–173.
[20] J. Talonen, H. Hänninen, Acta Mater. 55 (2007) 6108–6118. [27] G. Saller, K. Spiradek-Hahn, C. Schen, H. Clemens, Mater. Sci. Eng. A 427 (2006)
[21] T.S. Byun, Acta Mater. 51 (2003) 3063–3071. 246–254.
[22] E.H. Lee, T.S. Byun, J.D. Hunn, M.H. Yoo, K. Farrell, L.K. Mansur, Acta Mater. 49 [28] D. Khulmann-Wilsdorf, N.R. Comins, Mater. Sci. Eng. A 60 (1983) 7–24.
(2001) 3269–3276. [29] A. Kobel, J.D. Embury, M. Hatherly, P.M. Martin, H.W. Erbsloh, Acta Metall. 34
[23] R.D.K. Misra, B. Ravi Kumar, M. Somani, P. Karjalainen, Scripta Mater. 59 (2008) (1986) 1999–2009.
79–82. [30] A. Kobel, P. Martin, Acta Metall. 34 (1986) 1905–1909.
[24] D. Khulmann-Wilsdorf, Mater. Sci. Eng. A 315 (2001) 211–216. [31] J.C. Huang, G.T. Gray III, Acta Metall. 37 (1989) 3335–3347.
[25] H.-P. Karnthaler, B. Schügerl, Strength of Metals and Alloys, Pergamon Press,
Oxford, 1979, p. 205.

You might also like