Vous êtes sur la page 1sur 26

International Journal of Plasticity 25 (2009) 22222247

Contents lists available at ScienceDirect

International Journal of Plasticity


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

Experimental investigation on martensitic transformation


and fracture morphologies of austenitic stainless steel
Arpan Das *, Soumitra Tarafder
Fatigue & Fracture Group, Materials Science & Technology Division, National Metallurgical Laboratory (Council of
Scientic & Industrial Research) Jamshedpur 831 007, India

a r t i c l e

i n f o

Article history:
Received 24 May 2008
Received in nal revised form 17 March
2009
Available online 28 March 2009
Keywords:
Deformation induced martensite
Strain rate
Image processing
Dimple density
Dimple diameter

a b s t r a c t
Formation and nucleation mechanisms (i.e. c ? e, c ? a0 ,
c ? deformation twins ? e ? a0 and c ? e ? a0 ) of deformationinduced martensite (DIM) have been studied through analytical
transmission electron microscopy (TEM) after tensile deformation
of AISI 304LN stainless steel at various strain rates (SR) at room
temperature (RT). Quantitative metallography has been employed
extensively to assess martensitic transformation (MT) as function
of strain and SR. It has been observed that the enhancement of
SR during tensile deformation promotes the early formation of
DIM, while suppressing its saturation value at fracture. Fracture
surface morphologies and dimple geometries (i.e. dimple density,
dimple diameter and dimple size distribution) have been quantied through image processing (IP) of tensile fractographs. It is
noted that at lower SR, dimple density is high while dimple diameter is smaller, and vice versa. Concomitantly, the strength is noted
to be low and ductility is high at lower SR, and vice versa. DIM has
been found to be responsible for high dimple density at low SR. At
high SR, MT is suppressed and hence low dimple density. The
variation in SR dependent MT accounts for the variation in dimple
metrics vis--vis tensile properties.
2009 Elsevier Ltd. All rights reserved.

1. Introduction
Austenitic stainless steels of various types are potential candidate materials for use in chemical and
power generating industries for their excellent combination of mechanical properties, both at high
and low temperatures, and very good oxidation and corrosion resistance properties (Angel 1954; Lula,
* Corresponding author. Tel.: +91 657 2345225; fax: +91 657 2345213.
E-mail address: dasarpanl@yahoo.co.in (A. Das).
0749-6419/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2009.03.003

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2223

1966; Narutani et al., 1982; Ganesh Sundara Raman and Padmanabhan, 1994; Choi and Won, 1997). In
recent years, AISI 304LN stainless steel has evinced considerable interest among scientists and engineers because of its enhanced yield strength (YS), high work hardening characteristic, excellent weldability and improved intergranular corrosion resistance property (Lula, 1966; Narutani, 1989; Huang
et al., 1989; Ganesh Sundara Raman and Padmanabhan, 1994; Choi and Won, 1997; Xue et al., 2007).
Under ambient condition of loading, austenitic stainless steels commonly experience ductile failure
governed by dislocation ow, their mutual interactions and interactions with other second-phase particles/phases, etc. The mechanism of ductile failure is well established and known to be constituted of
three distinct events: void nucleation, growth and coalescence to form a crack, followed by nal fracture (Gurson, 1977; Tvergaard, 1981; Thomason, 1990; Benzerga et al., 2004; Chae and Koss, 2004;
Bandstra et al., 2004). Anderson (1995) concluded that the process of ductile fracture constituting
the stages of void nucleation and growth are predominantly inuenced by the nature of dislocationparticle and dislocationdislocation interactions, which in turn are governed by the state of
strain hardening in a material. It has been reiterated that voids do not necessarily nucleate from inclusions alone (Argon et al., 1975), but are also nucleated from precipitates (Zurek and Meyers, 1970;
Christy et al., 1986), phase constituents (Dieter, 1993; Anderson, 1995; Rivas et al., 2000; Erdogan
and Tekeli, 2002; Benzerga et al., 2004; Poruks et al., 2006; Das et al., 2006), grain boundaries (Barai
and Weng, 2008), shear bands (Tvergaard, 1981) or even at substructural heterogeneities such as dislocation pile-ups (Zurek and Meyers, 1970; Christy et al., 1986) in the metal matrix, and grow under
the inuence of favourable plastic strain and hydrostatic stress (Gurson 1977; Tvergaard, 1981, 1990;
Tvergaard and Needleman, 1984; Anderson 1995). The sub-structural alterations due to dislocation
interactions and microstructural changes that occur during the deformation process would inuence
not only the manner in which the voids nucleate but also the process through which they grow and
coalesce. Deformation paths characterised by SR, extent of deformation and temperature would play
an important role in controlling not only the mechanical behaviour of the material, but also void
nucleation and growth process.
The issue of void growth modelling is long standing and has received considerable attention for the
last 40 years. Theoretical work related to analytical and constitutive modelling of void growth include:
McClintock (1968), Rice and Tracey (1969), Gurson (1977), Tvergaard (1981, 1990); Tvergaard
and Needleman (1984); Koplik and Needleman (1988), Karr et al. (1989), Thomason (1990), Llorca
et al. (1991), Qiu and Weng (1993), Pardoen and Hutchinson (2000,2003), Liu et al. (2003), Wen
et al. (2005), Gan et al. (2006), Hammi and Horstemeyer (2007), Monchiet et al. (2008), Borg
et al. (2008), Zairi et al. (2008), Barai and Weng (2008), Quang and He (2008), Sanchez et al. (2008),
Hsu et al. (2009), etc.
There are also studies, summarised and cited below in Table 1, correlating fracture features with
mechanical properties. For instance, Goods and Brown (1979) concluded that there is a linear relationship between the measured dimple density on the fracture surface and the mean inclusion spacing for
a number of aluminium alloys. Broek, 1973 also found that void initiation strongly depends on size of
the inclusion particles at the nucleation sites. He demonstrated that there is a strong correlation between the dimple size and inclusion spacing. In some structural materials, for instance in high
strength steel (Cox and Low, 1974; Hancook and Mackenzie, 1976) and in aluminium alloy (Hahn
and Roseneld, 1975), failure results from nucleation, growth and coalescence of different sizes of
voids. Garrison and Moody (1987) noticed that if a material contains only one type of second-phase
particle, void initiation will not occur simultaneously at all of the particles. Typically, voids nucleate
at the larger particles rst and the choice is also inuenced by the shape of the particle, its orientation
(if not equiaxed), and by the strength of the particlematrix interface.
Since deformation and fracture are inuenced, to a large extent, by the same set of factors, a fracture surface would keep an imprint of the entire deformation process. This, however, requires careful analysis of the fracture feature morphologies and its relation to the mechanical properties. The
present authors (Das and Tarafder, 2008; Das et al., 2008a,b,c) have discussed the detailed philosophy
of the inter-relationship between deformation properties and fracture. Fig. 1 schematically represents
the basic concept of such inter-relationship.
In the present context, it is worthwhile to note that in some grades of austenitic stainless steels,
DIM forms during service depending upon austenite (fcc) phase stability (Jones and Bhadeshia,

2224

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Table 1
Review of published literatures correlating fracture features with deformation properties.
Researchers

System

Findings

Calhoun and Stoloff,


1970

Mg-alloy

The dimple size on fracture surface qualitatively agreed with the distance of the
small inclusions visible in surface replicas

Kirman, 1971

Al-alloy

A transition from transgranular dimple rupture to intergranular dimple rupture


with ageing time

Low et al., 1972

Al-alloy

The inter centre spacing of particles is comparable to dimple spacing

Cox and Low, 1972

AISI 4340 steel

There is a close correspondence between average spacing of large dimples with


the spacing of non-metallic inclusions

Broek, 1973

Al-alloy

There is a linear relationship between dimple density on fracture surface and


mean inclusion spacing

El-Soudani and
Pelloux, 1973

Al-alloy

Large dimple (i.e. spacing 10 lm) associated with large inclusions and small
dimples (i.e. spacing 0.7 lm) associated with intermediate particles

Speich and Spitzig,


1982

4340 steel

The inuence of inclusion volume fraction on the smooth axisymmetric and plane
strain tensile ductilities

Garrison and Moody,


1987

Low alloy steels

Low strength material exhibit very low tensile ductilities if the volume fraction of
a void nucleating second phase is sufciently high

Garrison and Moody,


1987

Ultra high
strength steels

The area of the ne dimpled region decreases as the plane strain ductility
increases

Karlik et al., 2000

Fe28Al4Cr
0.1Ce alloy

There is a combined inuence of deformation temperature and strain rate on the


fracture morphology (& mechanical properties)

Rizal and Homma,


2000

Al-alloy

The dominant dimple size depends on pulse duration

Erdogan and Tekeli,


2002

AISI 8620 steel

There is a change in fracture mode from ductile like depression mode to cleavage
type with increasing particle size

Pardoen et al., 2003

Ductile materials

Calculated the relationship between fracture toughness and microstructural


details for ductile materials based on a dilatational plasticity constitutive model

Wu and Kim, 2003

SA 508 steel

Dimple size becomes smaller and shallower for hydrogen charged specimens
compared to as received specimen

Xiao et al., 2003

Zr-based bulk
metallic glass

There is a gradual change in fracture surface morphology from cleavage veins to


microvoids coalescence dimples with increasing strain rate

Lee et al., 2004

304 L SS

Higher strain rate tends to reduce the size of the dimples and to increase their
density

Lee et al., 2006

Sintered 316 SS

There is a decrease in depth and density of voids on the fracture

Salemi and Abdollahzadah, 2008

NiCrMoV steel

The size of the dimples increased as the tempering temperature is increased

Lee et al., 2008

Low C steel

The depth of dimples increases with increasing strain rate

Lee et al., 2008

Medium C steel

The size of knobbing feature area increases with increasing strain rate

Miura et al., 2008

Carbon steel

The shape and distribution of voids on the fracture surface is found to be sensitive
to deformation temperature and strain rate

1997) under service or testing conditions (Olson and Cohen, 1972, 1976a,b; Tamura, 1982; Hecker
et al., 1982; Talonen, 2007). Stacking fault energy (SFE) plays an important role in determining
austenite phase stability, since it controls defect formation and is strongly dependent on chemistry
of the alloy and temperature (Talonen, 2007). Wang et al. (2005) predicted that austenite becomes
mechanically stable at a true strain of 2. Deformation of austenitic stainless steel results in the transformation of parent c (fcc) austenite to e (hcp) and a0 (bcc) martensites. The extent of DIM formation
depends upon several factors such as: material chemistry, temperature, plastic strain, SR, stress state,
deformation mode, grain size, grain orientation, etc. (Angel, 1954; Murr et al., 1982; Hecker et al.,
1982; Huang et al., 1989; Shrinivas et al., 1995; Talonen, 2007). Earlier research on AISI 304 stainless

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2225

MICROSTRUCTURE
Chemistry, Crystal
structure, Defects

Void-initiating
particles

Accumulation of
Stress and Strain

Correlations

Grains,
Grain-boundaries
Orientations

Mechanical
properties

Fracture
features

Fig. 1. Schematic representation of the concept of inter-relationship between deformation properties and fracture.

steel postulated that e (hcp) martensite is an intermediate or transient phase during: c (fcc) ? a0 (bcc)
transformation with an orientation relationship with parent austenite as: (1 1 1)c || (0 0 0 1)e, and that
the plates of a0 (bcc) martensite having {2 2 5}c habit plane nucleate from e (hcp) martensite
(Mangonon and Thomas, 1970; Breedis and Kaufman, 1971; Bracke et al. 2007). These ndings can
be rationalized considering the low SFE (20 mJ/mm2) of AISI 304 stainless steel (Talonen, 2007).
More recent observations with HR-TEM have demonstrated that both types of martensite can be produced independently during deformation, i.e. the presence of e (hcp) martensite is not a prerequisite
to a0 (bcc) martensite formation (Brooks et al., 1979; Shrinivas et al., 1995). While e (hcp) martensite
forms from overlapping stacking faults on every other {1 1 1}c plane by the passage of a=6h1 1  2i
Shockley partial dislocations; a0 (bcc) martensite is reported to form at intersection between two micro shear-bands (Venables, 1962; Kelly, 1965; Lecroisey and Pineau, 1972; Suzuki et al., 1977; Brooks
et al., 1979; Bracke et al., 2007). The shear bands can be e (hcp) martensite, mechanical twins, dense
stacking faults or twin boundaries. However during plastic deformation, the e (hcp) martensite eventually transforms to a0 (bcc) martensite at higher strains, so that at larger strains only a0 (bcc) martensite is observed. The mechanism proposed for this MT is the shifting of faulted planes in the e
(hcp) martensite phase by a second shear, a=12h1 1  2i, resulting in a a0 (bcc) martensite structure
(Kelly, 1965; Brooks et al., 1979; Bracke et al., 2007). This MT sequence directly inuences the
mechanical behaviour of the material. With the formation of DIM, the microstructure changes from
single phase austenite to a composite structure, with martensite behaving as a reinforcing phase,
thereby signicantly affecting the mechanical behaviour and enhancing the work hardening rate
(Suzuki et al., 1977; Spencer et al., 2004).
Several models of the diffusion-less phase transition of austenite into martensite have been proposed in the last 35 years. The constitutive modeling related to MT micro-mechanisms include the
works of Olson and Cohen (1972), Narutani et al. (1982), Narutani (1989), Leblond (1989), Stringfellow
et al. (1992), Tanaka et al. (1994), Diani et al. (1995), Tomita and Iwamoto (1995), Tanaka et al. (1996),
Jones and Bhadeshia (1997), Fischer et al. (1997), Levitas et al. (1998), Cherkaoui et al. (1998), Diani
and Parks (1998), Iwamoto et al. (1998), Stupkiewicz and Petryk (2002), Tang et al. (2002),
Mori et al. (2004), Han et al. (2004), Oberste-Brandenburg and Bruhns (2004), Idesman et al.
(2005), Serri et al. (2005a), Turteltaub and Suiker (2005), Garion et al. (2006), Bontcheva et al.
(2007), Hallberg et al. (2007), Dan et al. (2007), etc.
Experimental investigations to understand the tensile behaviour of metals and alloys, usually carried out by testing of smooth or notched tensile specimens at SRs ranging between 103 and 101 s1,
are generally regarded as isothermal processes and are also termed as quasi-static deformation. However, a critical assessment on the effect of SR spanning the quasi-static domain of experiments for tensile behaviour of materials where deformation-induced phase transformation occurs has rarely been
carried out. It is commonly known that with increase in SR, yield strength (YS) and ultimate tensile
strength (UTS) gradually increase with simultaneous lowering in ductility [i.e. elongation (%EL) and
reduction in area (%RA)]. It is thus expected that the combined inuence of SR and deformation induced phase transformation would lead to a phenomenal change in the plastic deformation process
and tensile properties.

2226

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

In the present investigation, an attempt has been made to study the effect of SR on the formation of
DIM in AISI 304LN stainless steel, its role in void formation on tensile straining at SRs ranging between
0.0001 and 1.0 s1, and quantitatively correlate fracture features (through dimple number density,
dimple diameter and its distribution) with SR dependent tensile properties. To meet the objective
of the present investigation, the following work has been carried out: (1) quantitative assessment
of the evolution of DIM as a function of strain for different SRs through quantitative metallography
(2) analytical TEM study to ascertain the types of DIM, their nucleation sites and to understand the
various MT mechanisms, and (3) quantitative measurements of fracture surface features (i.e. dimple
geometries) on tensile fracture surfaces using SEM imaging and IP technique.
2. Experimental
Nuclear grade AISI 304LN stainless steel has been used in the present investigation. The as-received
material was available in the form of pipe with 320 mm outer diameter and 25 mm wall thickness.
Chemical composition (in wt%) of the material is: C 0.03, Mn 1.78, Si 0.65, S 0.02, P 0.034, Ni 8.17,
Cr 18.73, Mo 0.26, Cu 0.29, N 0.08 and the balance Fe. The Md30 temperature of the material at which
50% of the austenite transforms to martensite at a true strain of 0.30, calculated from the equation of
Angel (1954), is found to be 2.756 C. Tensile tests were carried out on specimens with 6 mm gauge
diameter tted with a 25 mm gauge length extensometer at SRs of: 0.0001, 0.001, 0.01, 0.1 and
1.0 s1 at RT using a servo-hydraulic testing system (INSTRON 8862) in laboratory air. Few tensile
tests have also been carried out at 285 C at selected SR conditions. All tests were carried out under
computer control such that a minimum of 800 data points are collected for constructing the engineering stressstrain curve of the material. Tensile properties data at RT and at high temperature are tabulated in Table 2.
Fractured tensile specimens were longitudinally sectioned along the mid-plane, polished and
etched with a mixture of HCl and HNO3 in 2:1 ratio. Few drops of C2H5OH were used to reveal the
DIM using optical microscopy. Various techniques, like EBSD, XRD, IP, magnetic measurement etc.,
are commonly used to determine the amount of austenite transforming to martensite (Zhao et al.,
2001; Nagy et al., 2004; De et al., 2004). In this study, extensive IP has been carried out on optical
micrographs obtained from cross-sectional planes along the specimen axis and parallel to the fracture
surface for quantitative information on DIM in specimens tested at different SRs. The schematic of
specimen moulding is shown in Fig. 2. Starting from the fracture end of the specimens, a large number
of elds (i.e. several optical images) were sequentially viewed under an optical light microscope at
predetermined diametral locations, and images of the microstructure were digitally recorded with a
magnication of 400. DIM appears darker (i.e. grey) in the bright background of matrix austenite under the optical microscope. These optical images were analysed after appropriate grey-thresholding to
obtain the volume fraction of DIM (i.e. total e and a0 phases) as a function of true strain. This method of
estimating the amount of DIM was employed for all the tensile specimens fractured at different SRs.
The true strain corresponding to the amount of DIM that has been quantitatively determined was estimated by measuring the diameter (D) of the specimen at the specic transverse plane with the help of
a travelling microscope (ALMICRO) and making use of the relationship: True strain(e) = ln(Ao/
A) = 2ln(Do/D) (Dieter, 1993), where Ao is the original cross-sectional area based on original diameter
Do, and A is the cross-sectional area at selected transverse plane where the diameter is D. The average
Table 2
Comparison of tensile properties of AISI 304LN stainless steel at RT and at 285 C.
SR (s1)

1
0.1
0.01
0.001
0.0001

YS (MPa)

UTS (MPa)

%EL

%RA

RT

285 C

RT

285 C

RT

285 C

RT

285 C

464
418
382
353
340

235
240
252
222

734
688
675
671
680

511
509
536
518

36.22
46.24
53.66
69.62
70.61

44
52
50
32

66.88
82.2
85.84
88.2
87.64

74
69
66
69

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

d2

d3

2227

d
d4 d5 6

d1
Frame
Fracture
surface

Specimen

Mould

Loading direction

Fig. 2. Schematic of metallography sample inside the mould for quantifying DIM volume fraction at various strain level.

result obtained from all the images obtained at a particular D has been considered as the volume fraction of DIM formed in the cross-sectional plane of interest.
For investigating the nucleation sites for deformation-induced phase transformation (or plasticityinduced phase transformation) and to understand the mechanisms of evolution of deformationinduced MT, thin discs of 2.0 mm thickness were extracted from uniformly deformed regions of the
tensile specimens using a slow speed cutter (ISOMET 4000), and the discs were mechanically thinned
down to 0.10 mm thickness using increasingly ner grades of grit paper. After every stage of mechanical polishing, the specimens were examined under optical microscope in order to ensure that no
phase transformation has occurred during the thinning process (Das et al., 2008b,c). Following
mechanical polishing, 3.0 mm diameter discs were extracted from the thinned discs using a punching
apparatus, and subsequently thinned in a twin-jet electro polisher using 9:1 aceticperchloric acid
mixture at 45 V until perforation. The perforated thin foils were examined under a TEM (PHILIPS
CM 200) in a double tilt holder at an operating voltage of 200 kV.
Fracture surfaces were carefully examined under SEM (JEOL JSM 840A SEM) to record fractographic features. A set of elds were observed at an operating voltage of 20 kV throughout. A suitable
magnication (500) was used in all cases so that representative fracture features are recorded. The
dimple population and the microstructural features just adjacent to the fracture surfaces have also
been observed under SEM with secondary electron imaging mode.
Extensive IP technique was employed on the digital fractographs to characterise the two-dimensional geometry of dimples (i.e. dimple number density, dimple diameter and their distribution) on
the fracture surfaces using a commercial software (CLEMEX PE 3.5) ofine. The software had provision
for writing user dened program routines for quantitative analysis. A routine was developed incorporating several IP operations such as ltering, image enhancement, thresholding, grey level adjustment,
scale calibration, object identication, object measurements, etc., to get a quantitative estimation of
dimple geometries (i.e. diameter, density, etc.). All the fractographs from each test condition were analysed to arrive at an average value of the above parameters.
3. Result and discussions
3.1. Microstructure and tensile properties of AISI 304LN stainless steel
SEM micrograph shown in Fig. 3 reveals that in the as-received condition, the microstructure of the
investigated steel consists of polygonal grains of austenite with annealing twins, characteristic of
austenitic stainless steels, interspersed in some grains. The average grain size measured by the line
intercept method is found to be 70 lm. Engineering stressstrain curves of the steel tested at ve

2228

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Fig. 3. SEM microstructure of AISI 304LN stainless steel showing metastable austenitic grains and twins.

different SRs at RT are shown in Fig. 4. It is observed that in common with most metallic materials, the
ow stress increases with increase in SR. Fig. 5 reveals that while there is a steady increase of YS with
increase in SR, the effect on UTS is not very prominent and remains almost constant at an average
value of 678 MPa until SR exceeds 0.1 s1. The variation in ductility parameters shown in Fig. 6
similarly reveals that while %EL decreases steadily with increase in SR, %RA decreases very little until
the SR exceeds 0.1 s1. The ow curve of metals and alloys in the regime of uniform plastic
deformation can be expressed by r = Ken, where K is the strength coefcient and n is the strain
hardening exponent (Dieter, 1993). In the present study, it is found that n is strongly depending upon
SR because of variation in stress-induced phase transformation behaviour in this material. As shown in
Fig. 7, n decreases with increasing SR.
The increase in ow stress with increase in SR can be rationalised from the viewpoint of dislocation
activity during the process of plastic deformation. The initial dislocation density and dislocation congurations were constant, xed by the processing condition of the pipe from which the tensile specimens have been fabricated. The increase in ow stress with increase in SR is governed by the
 bqv, where, b is the Burgers vector, q is the mobile dislocation density and v is the
equation: 
average dislocation velocity (Dieter, 1993). To maintain the higher imposed SR, v needs to be increased
0
with simultaneous increase in ow stress according to the relationship: v Arm (Dieter, 1993).
800

Engineering stress, MPa

700
600
500
400
300

Strain rate (s-1)


0.0001
0.001
0.01
0.1
1

200
100
0
0

10

20

30

40

50

60

70

80

Engineering strain, %
Fig. 4. Engineering stressstrain curve of AISI 304LN stainless steel.

Strength properties, MPa

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

950
900
850
800
750
700
650
450
400
350
300
250
200
150

2229

YS
UTS
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 5. Variation of strength properties in AISI 304LN stainless steel with SR.

Ductility properties, %

95
90
85
80
70
60
50
%EL
%RA

40
30
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 6. Variation of ductility properties of AISI 304LN stainless steel with SR.

Strain hardening exponent, n

0.50
0.45
0.40
0.35
0.30
0.25
0.20
0.15
0.10
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 7. Variation of strain hardening exponent n as a function of SR.

Thus with increase in SR, YS steadily increases. After initial yielding, ow stress increases on continued
plastic deformation due to generation of dislocations and interaction of dislocations among themselves and with various types of microstructural barriers. Additionally, formation of DIM in metastable
austenitic stainless steel becomes another source of hardening (Hecker et al., 1982; Murr et al., 1982;

2230

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Huang et al., 1989; Talonen, 2007). Guimares and De Angelis (1974) have suggested that dislocation
production in the austenite/martensite interfaces causes accumulation of dislocations in the austenite
phase and accounts for the enhanced work hardening when DIM is formed in a material.
While the increase in YS with SR can be explained from the premise of enhanced ow stress to sustain higher dislocation velocities, the near constancy of UTS for SRs ranging between 0.0001 and
0.1 s1 has to be appreciated from the point of view of ow localisation originating from the initiation
of microvoids. In the present investigations, it has been observed, as discussed later, that void initiation is connected to DIM formation. It has been shown that DIM evolution is a function of the SR,
decreasing at increasing SRs. This is thought to inuence ow localisation through void initiation,
and consequently lead to the variation of UTS as observed. The formation of DIM during deformation
leads to the generation of a dynamically varying composite microstructure consisting of austenite matrix and DIM islands. Opposing inuence of higher ow stress at higher SRs and availability of void
initiation sites at DIM and other particulate constituents lead to the apparent insensitivity of the
UTS to SR until high values of the SR (1.0 s1). While work hardening during tensile deformation controls the ow behaviour, nucleation of microvoids (and their growth characteristics), governed by the
combined inuence of SR and the dynamic evolution of DIM, nally controls the load-bearing capacity
at the UTS.
3.2. DIM formation in AISI 304LN stainless steel at ambient temperature
Representative SEM image of the tensile deformed specimen in Fig. 8 shows the distribution of
martensite in austenite matrix. Such type of martensite is commonly known as DIM. It is revealed that
the DIM is not uniformly distributed throughout all the grains of the austenitic matrix. Such non-uniform distribution of martensite indicates the inhomogeneous nature of deformation in micro-scale.
Fig. 9 shows the effect of SR on the variation of DIM volume fraction as a function of true tensile strain.
It is observed that irrespective of SR, the nature of variation of DIM volume fraction with true strain is
similar (i.e. sigmoidal). In this regard, reports of early investigations (Hecker et al., 1982; Murr et al.,
1982; Lee and Lin, 2002; De et al., 2004; Lee et al., 2006; Lichtenfeld et al., 2006; Talonen, 2007) indicate both linear and non-linear variation of DIM volume fraction with the amount of deformation.
While Talonen, 2007 observed that martensite formation in AISI 301LN stainless steel follows a
non-linear relation with true strain, Lee and Lin (2000, 2004) and Lichtenfeld et al. (2006) observed
linear relationship in AISI 309 and 304L stainless steels. In this present investigation, the nature of
variation of martensite volume fraction with true strain follows similar pattern as that observed by
Hecker et al. (1982), De et al. (2004), Talonen (2007). A close look at Fig. 9 reveals that at lower strain

Fig. 8. SEM microstructure of deformed AISI 304LN stainless steel showing DIM morphologies.

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2231

Martensite volume fraction, %

35
30
25
20

Inflection point
Strain rate, s-1
1.0
0.1
0.01
0.001
0.0001

15
10
5

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3

True strain
Fig. 9. Volume fraction of DIM as a function of true strain at various SR.

levels, the amount of martensite increases very rapidly with true strain in a linearly proportional manner, after which the rate of transformation slows down. On comparing the curves for different SRs, it is
observed that while the increase in SR favours martensite formation at low strain level, the maximum
amount of martensite formed due to tensile deformation is reduced with increase in SR. This is mainly
attributed to the local variation of SFE of the material. At the higher SR, the heat of deformation is retained in the specimen and temperature increases during plastic deformation. Austenite stability increases because of this, and less martensite is formed. Bressanelli and Moskowitz (1966) clearly
demonstrated this effect in AISI 301 and 304 stainless steels by applying various SRs during tensile
testing. It is also noted that the SFE in austenitic stainless steels decreases as the temperature is decreased (Lecroisey and Pineau, 1972). Guntner and Reed (1962) proposed that the initial easy deformation stage is due to the e (hcp) martensite formation and the rapid hardening stage is due to the
evolution of a0 (bcc) martensite formation, since e (hcp) martensite tends to disappear at the
beginning of the rapid hardening stage (Mangonon and Thomas, 1970). Tamura (1982) interpreted
the rst stage as being caused by the normal slip deformation mode plus deformation due to e
(hcp) martensite formation. Petit et al. (2007) showed after certain amount of true strain (10%), volume fraction of a0 (bcc) martensite increases with increasing true strain whereas e (hcp) martensite
volume fraction decreases with increasing true strain for AISI 304 stainless steel. It may be noted that
on continued plastic deformation, piling up of dislocations against various types of barriers and
dislocationdislocation interactions can cause a local rise in temperature (Lichtenfeld et al., 2006;
Talonen, 2007). Austenitic stainless steel being poor in thermal conductivity, this local rise in temperature is expected to be high when the SR is increased. The inections (i.e. slope transition points) in
the curves (Fig. 9), indicating sluggish MT after some critical level of plastic deformation depending
upon the SR, and the reduced amount of martensite formation (shown in Fig. 10) with increasing
SR is attributed to this local rise in temperature. The inuence of adiabatic heating in decreasing
the MT has earlier been reported by Murr et al. (1982), Hecker et al. (1982), and Talonen (2007).
To understand the mechanism of DIM formation in AISI 304LN stainless steel, samples prepared
from the uniformly deformed regions of tensile specimens tested at various SRs were examined under
an analytical TEM. The mode of MT induced by plastic deformation depends on the stress state, temperature, strain, SR, and deformation methods. For instance, in AISI 304 stainless steel, the modes of: c
(fcc) ? e (hcp), c (fcc) ? a0 (bcc), c (fcc) ? e (hcp) ? a0 (bcc) and c (fcc) ? deformation twin ? a0
(bcc) were proposed to be the MT mechanisms for uniaxial tension (Lee and Lin, 2000; Lee and Lin,
2002; Talonen, 2007) and wire drawing (Choi and Won, 1997). Figs. 1117 show representative
TEM images corresponding to different SRs illustrating the various nucleation sites of DIM and the
multitude of MT mechanisms that occur. Corresponding selected area diffraction pattern (SADP) analyses, taken from the encircled region with the TEM micrographs, are inset in the gures. The study
reveals that there are various possible nucleation sites of martensite: shear-band intersections, isolated shear-band, shear-band grain boundary intersection, grain boundary triple point, etc. At SR of

2232

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Martensite volume fraction, %

28
26
24

Strain rate, s-1


1.0
0.1
0.01
0.001
0.0001

22
20
18
0.25

0.30

0.35

0.40

0.45

0.50

0.55

0.60

True strain
Fig. 10. Inexion point variation (% DIM) with SR.

Fig. 11. TEM bright eld image showing the formation of a0 -martensite (bcc) with zone axis of 1 1 3 in c-austenite (fcc) with
zone axis of 2 3 3 at the micro shear-band intersection at SR of 1.0 s1.

Fig. 12. TEM bright eld image showing the evolution of e-martensite (hcp) and a0 -martensite (bcc) with zone axis of
7 2  5 3 and 1 2 3, respectively, in c-austenite (fcc) matrix of zone axis [0 0 1] from isolated micro-shear band at SR of
0.001 s1.

1.0 s1 the a0 embryo nucleated at the shear-band intersection (Fig. 11) is a commonly known martensite nucleation site and it is referred as strain induced martensite (SIM) by Olson and Cohen
(1972, 1976a, b), Lecroisey and Pineau (1972) and Maxwell et al. (1974a,b). Lecroisey and Pineau

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2233

Fig. 13. TEM bright eld image showing the formation of e-martensite (hcp) and a0 -martensite (bcc) with zone axis of
0 1  1 1 and 1 1 1 respectively in austenite (fcc) matrix of zone axis [-1 1 1] at shear band-grain boundary intersection at SR
of 0.0001 s1.

Fig. 14. TEM bright eld micrograph showing the formation of a0 -martensite (bcc) with zone axis [0 1 2] in the matrix of caustenite (fcc) with zone axis of 1 1 3 at grain boundary triple point at SR of 0.0001 s1.

Fig. 15. TEM bright eld micrograph showing intersection of shear bands forming a0 (bcc) martensite with zone axis of [0 1 3]
and e (hcp) martensite with zone axis of 7 2 5  3 in c (fcc) austenite matrix of zone axis [0 1 3] during tensile deformation at
SR of 0.001 s1.

(1972), Maxwell et al. (1974a,b), Tamura (1982) and Petit et al. (2007) distinguished between stress
assisted and strain induced MTs (i.e. stress assisted martensite (SAM) and SIM) according to their

2234

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Fig. 16. (a) TEM bright eld image showing the formation of e-martensite (hcp) and a0 -martensite (bcc) with zone axis of
0 1  1 1 and 0  1  3 respectively in c-austenite (fcc) matrix with zone axis of 1  2  3 on the parallel micro shear bands
at SR of 0.01 s1 (b) SAD analysis and (c) Dark eld image.

nucleation sites. In Fig. 12 (SR of 0.001 s1), the evolving phase has been identied as a0 (bcc) and e
(hcp) martensites with zone axes of 1 2 3 and 7 2  5 3, respectively, both nucleated in the

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2235

Fig. 17. (a) TEM bright eld image showing the formation of a0 -martensite (bcc) with zone axis 0  1  3 in c-austenite (fcc)
matrix with zone axis of 1 1 3 on the parallel micro shear bands at SR of 0.001 s1 (b) SAD analysis and (c) Dark eld image.

2236

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

austenitic matrix (fcc) with zone axis [0 0 1] as indicated in the SADP analysis on the isolated micro
shear-band. The orientation relationships of Fig. 12 are found to be: 1 1  1c k0 1 1a0 and
1 1 0c k1 1  1a0 , which matches with variant V22 of KurdjumovSachs (KS) relationship (Murr
et al., 1982). Grain boundaryshear-band intersection shown in Fig. 13 (SR of 0.0001 s1) contains
both a0 (bcc) and e (hcp) martensite with zone axes of 1 1 1 and 0 1  1 1, respectively, in c (fcc)
austenite matrix with zone axis 1 1 1. The well-known model for MT: c (fcc) ? e (hcp) ? a0 (bcc)
has been proposed by Mangonon and Thomas (1970) and Maxwell et al. (1974a,b). According to the
study of Staudhammer et al. (1983), this type of transformation is known to be a stress-assisted transformation. Staudhammer et al. (1983) also concluded that SR does not affect the morphology of a0
(bcc) martensite. Alternatively, it has been proposed that the a0 (bcc) martensite forms by a direct
c (fcc) ? a0 (bcc) transition and that the direct transformation is merely stimulated by any e (hcp)
martensite that may have preceded it (Breedis and Kaufman, 1971; Maxwell et al., 1974a,b). Lecroisey
and Pineau (1972) proposed a direct c (fcc) ? a0 (bcc) dislocation interaction mechanism for the formation of SIM in deformed high Cr alloys. Fig. 14 (SR of 0.0001 s1) shows that a0 (bcc) martensite
with zone axis [0 1 2] nucleate at a grain boundary triple point in c (fcc) austenite matrix with zone
axis of [1 1 3]. Fig. 15 (SR of 0.001 s1) shows a number of micro shear-bands intersecting each other,
forming a0 (bcc) martensite with zone axis of [0 1 3] and e (hcp) martensite with zone axis of
[7 2 5 3] in c (fcc) austenite matrix of zone axis [0 1 3] during tensile deformation. A number of
parallel aligned micro shear-bands are shown in Fig. 16(a) and (b) (SR of 0.01 s1) generating e
(hcp) martensite with zone axis of [0 1 1 1] and a0 (bcc) martensite of zone axis [0 1 3] in c
(fcc) austenite matrix with zone axis [1 2 3]. The corresponding dark eld image is shown in
Fig. 16 (c). At SR of 0.001 s1, parallel and isolated micro shear-bands are shown in Fig. 17 (a), which
produce only a0 (bcc) martensite with zone axis [0 1 3] in c (fcc) austenite matrix of zone axis
[1 1 3] (Fig. 17b). The corresponding dark eld image is shown in Fig. 17(c).
TEM results in the present study show that the nucleation of martensite in AISI 304LN stainless
steel can simultaneously occur at locations other than intersecting shear-bands which is commonly
referred in the literature, and multiple mechanisms such as: c (fcc) ? e (hcp), c (fcc) ? a0 (bcc), c
(fcc) ? e (hcp) ? a0 (bcc) and c (fcc) ? mechanical twins ? a0 (bcc) can co-exist. It is also evident
from the TEM bright eld and dark eld images that in association with formation of DIM, large numbers of dislocations are generated. The various mechanisms for deformation-induced MT observed in
AISI 304LN stainless steel in the present investigation is listed in Table 3.
3.3. DIM vis--vis void formation
The micromechanism of ductile fracture involves nucleation of microvoids, growth and their coalescence. Microvoid formation occurs either by decohesion of second-phase particles from the matrix
or by cracking of the second-phase particles (Gurson, 1977; Garrison and Moody, 1987; Anderson,
1995; Benzerga et al., 2004; Chae and Koss, 2004; Bandstra et al., 2004). Assuming that the inclusion
volume fraction is constant and at a very low level, the main nucleation sites for void formation are
expected to be the interfaces between austenite and martensite formed during tensile tests. Any
variation in the tensile fracture morphology of the specimens tested at different SRs is thus attributed
to the differences in the amount of SR dependent DIM. A quantitative comparison reveals that dimple
Table 3
Nucleation sites for martensite formation with their zone axis calculated from SADP analysis in the present study.
Features (strain rates)

Phases
1

Shear-band intersection (1.0 s )


Isolated shear band (0.001 s1)
Shear band-grain boundary intersection (0.0001 s1)
Grain boundary triple point (0.0001 s1)
Shear-band intersections (0.001 s1)
Parallel micro shear-bands (0.01 s1)
Parallel micro shear-bands (0.001 s1)
*

a,c
e, a0 , c
e, a0 , c
a0 , c
e, a0 , c
e, a0 , c
a0 , c

SAM, stress assisted martensite; SIM, strain induced martensite.

Zone axis

Type (SIM/SAM*)

1 1 3 and 2 3 3
7 2  5 3, 1 2 3 and [0 0 1]
0 1  1 1, 1 1 1 and 1 1 1
[0 1 2] and 1 1 3
7 2 5  3, [0 1 3] and [0 1 3]
0 1  1 1, 0  1  3 and 1  2  3
0  1  3 and 1 1 3

SIM
SAM
SAM
SIM
SAM
SAM
SIM

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2237

number density is much less for straining of 1.0 s1 in comparison to that at 0.0001 s1. Lower dimple
number density associated with increased SR correlates well with the reduced tensile ductility observed in this study.
Representative SEM image of the sectioned tensile fracture specimen shown in Fig. 18 clearly reveals the sites for void nucleation. It is observed in Fig. 18 that void nucleation occurs at the interfaces
perpendicular to the tensile axis in consistence with the ndings of Poruks et al. (2006) for low carbon
bainitic steel and Erdogan and Tekeli (2002) for AISI 8620 steel. Microvoid formation associated with
martensite resulted from either decohesion at the austenitemartensite interface or from the separation of adjacent islands and localised deformation of martensite (Fig. 18). As discussed above, with the
lowering of SR, DIM volume fraction increases. As a result, at lower SRs, void nucleating sites increase
with the increase of austenitemartensite interfaces. In the present study, it is observed that martensite cracking in the specimen tested at 0.0001 s1 is less frequent and dimples are smaller in size in
comparison to the specimen tested at 1.0 s1. High density of microvoids near to the fracture surface
is generally associated with high plastic true strain in this region and under localised necking condition, dimple number density is known to increase towards the fracture surface according to Poruks
et al., 2006. Fractographic evidence clearly shows that the dimple number density is more in
0.0001 s1 (Fig. 19a) than in 1.0 s1 (Fig. 19e). There are a greater number of martensite particles
per unit volume and, therefore, more microvoid nucleation sites in specimen 0.0001 s1.
To conrm the association of microvoids with DIM, tensile tests were conducted at 285 C at
similar SRs. The tensile properties are tabulated in Table 2. There was no evidence of martensite found
in the deformed material microstructure under SEM and TEM for this high temperature.
Thermodynamically the formation of DIM is not feasible at this testing temperature. The fractographic
features of elevated temperature tensile fracture surfaces at selected SRs are shown in Fig. 20(a)(b). It
can be seen from these gures that at elevated temperature, even when the SR is varied from 0.0001 to
0.1 s1, the fractographic features are not altered signicantly. A comparison of this fractographic features with those at similar SRs in Fig. 19(a and d, respectively) immediately reveals the fact that a
majority of ne void network observed at RT deformation has disappeared when the test is performed
at elevated temperature. This further conrms that DIM act as the nucleation sites for smaller voids,
which gradually depletes when the MT is suppressed at high SRs (or at elevated temperature).
3.4. Void characterisation by IP and correlation with tensile properties
IP is a powerful tool for characterising the void morphologies (Deshpande et al., 2004) on fracture
surfaces. Voids within a fracture surface can easily be detected by IP due to the high degree of contrast

Fig. 18. Nucleation of void from martensite.

2238

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Fig. 19. SEM fractographs of ductile fracture feature of AISI 304LN stainless steel at SRs of (a) 0.0001 (b) 0.001 (c) 0.01 (d) 0.1
and (e) 1.0 s1 tested at ambient temperature. Void network as predicted by image processing on the SEM image at SR s of (f)
0.0001 (g) 0.001 (h) 0.01 and (i) 0.1 and (j) 1.0 s1. Close correspondence between the fractograph and the IP prediction may be
noted.

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2239

Fig. 20. Tensile fracture of AISI 304LN at two different SRs (i.e. (a) 0.0001 and (b) 0.1 s1) tested at 285 C. Difference in the
corresponding fractographic features at RT may be noted.

between the dark voids and the more reective voids at its peripheries. Banerjee (1988) has shown
how quantitative methods permit detection of subtle changes in the fracture surface topology inuenced by the microstructure of a material. He has also explained how the prole roughness parameter
and fractal can be attributed to the mechanical properties of 4340 steel at different heat treatment
conditions. The reliability of these methods has been statistically tested to give 95% condence level
(Deshpande et al., 2004). In the present study, dimple geometries on the tensile fracture surfaces of
specimens tested at RT at different SRs have been characterised in terms of dimple number density,
dimple size and their distribution.
SEM fractographs shown in Fig. 19(a)(e) reveal that irrespective of SR, all specimens have failed
through microvoid coalescence that is typical of ductile fracture. However, a close look at the fractographs reveals that there is signicant variation in dimple morphology arising due to the variation in
SR. The corresponding image processed dimple-networks shown in Fig. 19(f)(j) clearly elucidate that
depending upon SR, the mean size of dimples constituting the nal fracture changes. It is interesting to
visualise that with lowering of SR, the network of ne dimples becomes predominant, and with increase in SR, the network of ne dimples is gradually replaced by larger dimples. Fig. 21 showing
the distribution of dimple size for all the SRs studied, revealing that for SRs of 0.0001, 0.001 and
0.01 s1 the maximum population of dimples is of 0.5 lm mean diameter. With further increase
of SR, the distribution of all dimple size almost remains constant. The most frequent dimple size
was noted to be approximately between 0.35 and 0.65 lm range. The variation of density of smaller

Strain rate, s-1


1.0
0.1
0.01
0.001
0.0001

1250

Occurance

1000
750
500
250
0
0.0

0.5

1.0

1.5

2.0

Dimple diameter, m
Fig. 21. Dimple size distribution at various SR.

2240

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Smaller dimple (0.35-0.65 m)/unit frame, number/m2

dimples in the range of 0.350.65 lm with SR is shown in Fig. 22. It can be seen from this gure that
the dimples of the above dimension proportionally decrease as the SR is increased. This gure thus
conrms the earlier visual observation that the population of the ner dimples indeed decreases with
increasing SR. Fig. 23 shows that as the SR decreases mean dimple diameter decreases with simultaneous increase in dimple density.
It is known that formation of large number of small sized dimples increases both strength and ductility. Accordingly, it is a common practice to modulate the microstructure through variation in processing conditions so that nal fracture occurs with formation of large population of small sized
dimples for optimum strength-ductility combination. In the present study, the initial microstructure
of the steel, which is composed of austenite grains, is gradually changed to a composite type structure
composed of austenite and varying proportion of DIM depending upon SR. In such a composite aggregate, the interfaces between austenite and martensites act as void nucleation sites. It is thus logical to
consider that with increase in SR, the number of such void nucleating interfaces would decrease due to
the formation of lower amount of DIM and the growth opportunity of these smaller numbers of voids
would increase. As such with the increase in SR, voids would grow to a large extent before coalescing.
It is indeed observed as shown in the gures presented that with increase in SR, the mean dimple size
increases with simultaneous lowering of dimple density as shown in Fig. 23.

0.0014
0.0012
0.0010
0.0008
0.0006
0.0004
0.0002
0.0000
1E-4

1E-3

0.01

0.1

Strain rate, s-1

Dimple number density, number/m2

Fig. 22. Variation of density of smaller dimples with SR.

0.16

Density = 0.585 (Diameter)-1.88

0.14
0.12
0.10
0.08
0.06
0.04
0.02
0.00
2

Average circular diameter of dimple, m


Fig. 23. Variation of average circular diameter of dimples and dimple number density with respect to SR.

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2241

In the present investigation, dimple geometries that have been extracted by IP of the tensile fracture surfaces correlate well with the strength and ductility properties of the specimens tested at different SRs shown in Figs. 2428. It is observed that with increase in SR, while the variation of YS and
UTS follow similar (increasing) pattern as that observed for mean dimple diameter (Figs. 24 and 25),
the ductility, as measured by %EL and %RA, follow an inverse trend, showing that with increase in SR,
ductility property decreases (Figs. 24 and 26). Further, it is observed that the decrease in dimple density with increase in SR leads to simultaneous variation in strength and ductility property in such a
manner that with lowering of dimple density, YS and UTS increase with consequent lowering of ductility property as shown in Figs. 27 and 28, respectively.
DIM plays an important role in controlling void nucleation and growth, and hence the mechanical
behaviour. Formation of higher fractions of martensite promotes dislocation-generating sources,
which in turn alters the mechanical response. Martensite also provides more voids nucleating sites.
These two aspects synergistically control the manner in which the mechanical properties are altered
with SR. At low SR conditions, the availability of more martensite enhances the dislocation dynamics
and the formation of profuse (smaller) dimples. The strength under such conditions was therefore low.
At higher SR, suppression of MT decreased the dislocation generating sources and the availability of
(smaller) void nucleation sites. Therefore, the strength increased at the expense of ductility.

Dimple diameter, m

14
12
10
8
6
4
2
0
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 24. Variation of dimple diameter with respect to SR.

YS
UTS
Dimple diameter

730

475

720

450

710

425

400

375

350

325

700
690
680
670

Average circular diameter of dimple, m

500

YS, MPa

UTS, MPa

740

300
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 25. Variation of strength properties with average circular diameter of dimples at different SR.

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

90

75

70

65

60

55

50

80

EL, %

RA, %

85

75

45
70

%EL
%RA
Dimple diameter

40
65

Average circular diameter of dimple, m

2242

35
1E-4

1E-3

0.01

Strain rate,

0.1

s-1

Fig. 26. Variation of ductility properties with average circular diameter of dimples at various SR.

720

440

700

YS, MPa

460

UTS, MPa

730

710

0.00

480
YS
UTS
Dimple density

0.02
0.04
0.06

420

0.08

400

0.10

690

380

680

360

670

340

0.12
0.14

Dimple number density, number/m2

740

0.16
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 27. Variation of strength properties with dimple number density at various SR.

4. Conclusions
The conclusions from the present study can be summarised as follows:
1. When the SR is increased, the total volume fraction of martensite transformed decreased. This is
attributed to the rise in temperature at high SR which would suppress the MT. Moreover, an early
transformation has been observed at higher SR than at lower SR.
2. TEM results showed that martensite nucleates at many microstructural locations. They have been
identied at shear-band intersection, isolated shear-band, shear bandgrain boundary intersection,
grain boundary triple points, etc., mainly at defects within the microstructure.
3. Multiple MT mechanisms viz: c (fcc) ? e (hcp), c (fcc) ? a0 (bcc), c (fcc) ? mechanical twins ? a0
(bcc) and c (fcc) ? e (hcp) ? a0 (bcc) have been observed during deformation of AISI 304LN stainless steel at RT.

90
88
86
84
80
78
76

EL, %

RA, %

82

75

0.16

70

0.14

65

0.12

60

0.10

55

0.08

50

0.06

74
72
70

45

0.04
%EL
%RA
Dimple density

40

68
66

35

0.02

2243

Dimple number density, Number/m2

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

0.00
1E-4

1E-3

0.01

0.1

Strain rate, s-1


Fig. 28. Variation of ductility properties with dimple number density at various SR.

4. Fracture surface morphologies/geometries contain imprints of the deformation processes that


were operative.
5. Dimples of smaller size range primarily initiate from DIM.
6. At lower SR, dimple density was high and the mean dimple size was low. Dimple density was proportional to the inverse square of mean dimple diameter, as the SR is varied.
7. At low SR, both martensite and (small) dimple density were high, thus the ductility was high and
the strength was low.
8. At high SR, since MT is suppressed, dimple density was low and, therefore, the strength was high at
the expense of ductility.

Acknowledgements
The authors acknowledge the support and encouragement of Prof. S.P. Mehrotra, Director, NML and
Dr. R.N. Ghosh, previous Head of Materials Science & Technology Division. The authors express their
sincere gratitude to Dr. P.C. Chakraborti, Professor, Metallurgical & Materials Engineering Department,
Jadavpur University, for fruitful discussions and advice. Dr. S. Sivaprasad, Scientist, NML, is acknowledged for his help in mechanical characterisation. The authors also express their gratitude to
Dr. M. Ghosh and Dr. S. K. Das, both Scientists at NML, for their valuable advice and suggestions while
carrying out TEM and SEM characterisation.
References
Anderson, T.L., 1995. Fracture Mechanics: Fundamentals and Applications, third ed. CRC Press, Boca Raton, New York. 265.
Angel, T., 1954. Formation of martensite in austenitic stainless steels. Effects of deformation, temperature and composition.
Journal of Iron and Steel Institute 177, 165174.
Argon, A.S., Im, J., Safoglu, R., 1975. Cavity formation for inclusions in ductile fracture. Metallurgical Transaction A 6, 825837.
Bandstra, J.P., Koss, D.A., Geltmacher, A., Matic, P., Everett, R.K., 2004. Modelling void coalescence during ductile fracture of steel.
Materials Science and Engineering A 366, 269281.
Banerjee, K., 1988. Quantitative fractography: a modern perspective. Metallurgical Transactions A 19, 961971.
Barai, P., Weng, G.J., 2008. The competition of grain size and porosity in the viscoplastic response of nanocrystalline solids.
International Journal of Plasticity 24, 13801410.
Benzerga, A.A., Besson, J., Pineau, A., 2004. Anisotropic ductile fracture, Part I: Experiments. Acta Materialia 52, 46234638.
Bontcheva, N., Petrov, P., Petzov, G., Parashkevova, L., 2007. Finite element simulation of strain induced austenitemartensite
transformation and ne grain production in stainless steel. Computational Materials Science 40 (1), 90100.
Borg, U., Niordson, C.F., Kysar, J.W., 2008. Size effect on void growth in single crystals with distributed voids. International
Journal of Plasticity 24, 688701.

2244

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Bracke, B., Kestensa, L., Penninga, J., 2007. Transformation mechanism of a/ martensite in an austenitic FeMnCN alloy.
Scripta Materialia 57, 385388.
Breedis, J.F., Kaufman, L., 1971. Formation of HCP and BCC Phases in Austenitic Iron Alloys. Metallurgical Transactions A 2 (9),
23592371.
Bressanelli, J.P., Moskowitz, A., 1966. Effects of strain rate, temperature, and composition on tensile properties of metastable
austenitic stainless steels. Transactions of the ASM 59 (2), 223239.
Broek, D., 1973. The role of inclusions in ductile fracture and fracture toughness. Engineering Fracture Mechanics 5, 5560.
Brooks, J.W., Loretto, M.H., Smallman, R.E., 1979. B. Direct observations of martensite nuclei in stainless steel. Acta Metallurgica
27 (12), 18391847.
Calhoun, C.D., Stoloff, N.S., 1970. The effect of particles on fracture processes in Magnesium alloys. Metallurgical Transactions A
1, 997.
Chae, D., Koss, D.A., 2004. Damage accumulation and failure of HSLA-100 steel. Materials Science and Engineering A 366, 299
309.
Cherkaoui, M., Berveiller, M., Sabar, H., 1998. Micromechanical modeling of martensitic transformation induced plasticity (TRIP)
in austenitic single crystals. International Journal of Plasticity 14 (7), 597626.
Choi, J.Y., Won, J., 1997. Strain induced martensite formation and its effect on strain hardening behavior in the cold drawn 304
austenitic stainless steels. Scripta Materialia 36 (1), 99104.
Christy, S., Pak, H.R., Meyers, M.A., 1986. In: Murr, L.E., Staudhammer, K.P., Meyers, M.A. (Eds.), Metallurgical Applications of
Shock-wave and High Strain-rate Phenomena. Marcel Dekker, New York, NY, pp. 835863.
Cox, T.B., Low, J.R., Jr., NASA Tech. Rep. No. 4 on Research Grant NGR-39-087-003, Carnegie-Mellon Univ., 1972.
Cox, T.B., Low, J.R., 1974. An investigation of the plastic fracture of AISI 4340 and 18 Nickel-200 Grade Maraging steels.
Metallurgical Transactions A 5, 14571470.
Karr, Dale G, Kyungsik, C., 1989. A three dimensional constitutive damage model for polycrystalline ice. Mechanics of Materials
8, 5566.
Dan, W.J., Zhang, W.G., Li, S.H., Lin, Z.Q., 2007. A model for strain-induced martensitic transformation of TRIP steel with strain
rate. Computational Materials Science 40 (1), 101107.
Das, A., Tarafder, S., 2008. Geometry of dimples and its correlation with mechanical properties in austenitic stainless steel.
Scripta Materialia 59 (9), 10141017.
Das, A., Das, S.K., Sivaprasad, S., Tarafder, S., 2008a. Fractureproperty correlation in copper strengthened high strength low
alloy steel. Scripta Materialia 59 (7), 681683.
Das, A., Sivaprasad, S., Chakraborty, P.C., Tarafder, S., 2008b. Correspondence of fracture surface features with mechanical
properties in 304LN stainless steel. Materials Science and Engineering A 496 (12), 98105.
Das, A., Sivaprasad, S., Ghosh, M., Chakraborti, P.C., Tarafder, S., 2008c. Morphologies and characteristics of deformation induced
martensite during tensile deformation of 304LN stainless steel. Materials Science and Engineering A 486 (12), 283286.
Das, S.K., Chatterjee, S., Tarafder, S., 2006. Tensile fracture behaviour of microstructurally engineered Cu bearing high strength
low alloy steel. Materials Science and Technology 22 (12), 14091414.
De, A.K., Murdock, D.C., Mataya, M.C., Speer, J.G., Matlock, D.K., 2004. Quantitative measurement of deformation-induced
martensite in 304 stainless steel by X-ray diffraction. Scripta Materialia 50, 14451449.
Deshpande, S., Kulkarni, A., Sampath, S., Herman, H., 2004. Application of image analysis for characterization of porosity in
thermal spray coatings and correlation with small angle neutron scattering. Surface and Coating Tech 187, 616.
Diani, J.M., Sabar, H., Berveiller, M., 1995. Micromechanical modelling of the transformation induced plasticity (TRIP)
phenomenon in steels. International Journal of Engineering Sciences 33 (13), 19211934.
Diani, J.M., Parks, D.M., 1998. Effects of strain state on the kinetics of strain induced martensite in steels. Journal of Mechanics
and Physics of Solids 46 (9), 16131635.
Dieter, G.E., 1993. Mechanical Metallurgy, SI Metric Edition. Springer, Berlin, 145 pp.
El-Soudani, S.M., Pelloux, R.M., 1973. Inuence of inclusion content on fatigue crack. Metallurgical Transactions B 4 (2), 519
531.
Erdogan, M., Tekeli, S., 2002. The effect of martensite particle size on tensile fracture of surface-carburised AISI 8620 steel with
dual phase core microstructure. Materials and Design 23, 597604.
Fischer, F.D., Oberaigner, E.R., Tanaka, K., 1997. A micromechanical approach to constitutive equations for phase changing
materials. Computational Materials Science 9, 5663.
Gan, Y.X., Kysar, J.W., Morse, T.L., 2006. Cylindrical void in a rigid-ideally plastic single crystal, II. Experiments and simulations.
International Journal of Plasticity 22, 3972.
Ganesh Sundara Raman, S., Padmanabhan, K.A., 1994. Inuence of martensite formation and grain size on room temperature
low cycle fatigue behaviour of AISI 304LN austenitic stainless steel. Materials Science and Technology 10, 614620.
Garion, C., Skoczen, B., Sgobba, S., 2006. Constitutive modelling and identication of parameters of the plastic strain-induced
martensitic transformation in 316L stainless steel at cryogenic temperatures. International Journal of Plasticity 22, 1234
1264.
Garrison Jr., W.M., Moody, N.R., 1987. Ductile fracture. Journal of Physics and Chemistry of Solids 48 (11), 10351074.
Goods, S.H., Brown, L.M., 1979. The nucleation of cavities by plastic deformation. Acta Metallurgica 27, 115.
Guimares, J.R.C., De Angelis, R.J., 1974. Hardening by a deformation induced phase transformation. Materials Science and
Engineering A 15 (23), 291294.
Guntner, C.J., Reed, R.P., 1962. The effect of experimental variables including the martensitic transformation on the lowtemperature mechanical properties of austenitic stainless steels. Transaction of the ASM 55 (3), 399419.
Gurson, A.L., 1977. Continuum theory of ductile rupture by void nucleation and growth: Part I Yield criteria and ow rules for
porous ductile media. ASME Journal of Engineering Materials and Technology 99, 115.
Han, H.N., Lee, C.G., Oh, C.-S., Lee, T.-H., Kim, S.-J., 2004. A Model for deformation behavior and mechanically induced martensitic
transformation of metastable austenitic steel. Acta Materialia 52 (17), 52035214.
Hahn, G.T., Roseneld, A.R., 1975. Metallurgical factors affecting fracture toughness of aluminum alloys. Metallurgical
Transactions A 6, 653668.

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2245

Hallberg, H., Hakansson, P., Ristinmaa, M., 2007. A constitutive model for the formation of martensite in austenitic steels under
large strain plasticity. International Journal of Plasticity 23, 12131239.
Hancook, G.G., Mackenzie, A.C., 1976. On the mechanism of ductile failure in high strength steels subjected to multiaxial stress
states. Journal of Mechanics and Physics of Solids 24, 147169.
Hammi, Y., Horstemeyer, M.F., 2007. A physically motivated anisotropic tensorial Representation of damage with separable
functions for void nucleation, growth, and coalescence. International Journal of Plasticity 23, 16411678.
Hecker, S.S., Stout, M.G., Staudhammer, K.P., Smith, J.L., 1982. Effect of strain state and strain rate on deformation-induced
transformation in 304 stainless steel: Part I. Magnetic measurement and mechanical behaviour. Metallurgical Transactions
A 13, 619626.
Hsu, C.Y., Lee, B.J., Mear, M.E., 2009. Constitutive models for power-law viscous solids containing spherical voids. International
Journal of Plasticity 25 (1), 134160.
Huang, G.L., Matlock, D.K., Krauss, G., 1989. Martensite formation, strain rate sensitivity, and deformation behaviour of type 304
stainless steel sheet. Metallurgical Transactions A 20, 12391246.
Idesman, A.V., Levitas, V.I., Preston, D.L., Cho, J.-Y., 2005. Finite element simulations of martensitic phase transitions and
microstructures based on a strain softening model. Journal of the Mechanics and Physics of Solids 53, 495523.
Iwamoto, T., Tsuta, T., Tomita, Y., 1998. Investigation on deformation mode dependence of strain-induced martensitic
transformation in TRIP steels and modelling of transformation kinetics. International Journal of Mechanical Science 40 (2-3),
173182.
Jones, S.J., Bhadeshia, H.K.D.H., 1997. Kinetics of the simultaneous decomposition of austenite into several transformation
products. Acta Materialia 45, 29112920.
Karlik, M., Kratochvil, P., Janecek, M., Siegl, J., Vodickova, V., 2000. Tensile deformation and fracture micromorphology of an Fe
28Al4Cr0.1Ce alloy. Materials Science and Engineering A 289, 182188.
Kelly, P.M., 1965. The martensite transformation in steels with low stacking fault energy. Acta Metallurgica 13 (6), 635646.
Koplik, J., Needleman, A., 1988. Void growth and coalescence in porous plastic solids. International Journal of Solids and
Structures 24, 835853.
Kirman, I., 1971. Relation between microstructure and toughness in 7075 aluminium alloys. Metallurgical Transactions A 2,
17611769.
Leblond, J.B., 1989. Mathematical modelling of transformation plasticity in steels, II: Coupling with strain hardening
phenomena. International Journal of Plasticity 5, 573591.
Lecroisey, F., Pineau, A., 1972. Martensitic transformations induced by plastic deformation in the FeNiCrC system.
Metallurgical Transactions A 3 (2), 387396.
Lee, W.S., Lin, C.F., 2000. The morphologies and characteristics of impact-induced martensite in 304L stainless steel. Scripta
Materialia 43 (8), 777782.
Lee, W.S., Lin, C.F., 2002. Comparative study of the impact response and microstructure of 304L stainless steel with and without
prestrain. Metallurgical and Materials Transactions A 33 (9), 28012810.
Lee, W.S., Chen, J.I., Lin, C.F., 2004. Deformation and failure response of 304L stainless steel SMAW joint under dynamic shear
loading. Materials Science and Engineering A 381, 206215.
Lee, W.S., Lin, C.F., Liu, T.J., 2006. Strain rate dependence of impact properties of sintered 316L stainless steel. Journal of Nuclear
Materials 359, 247257.
Lee, W.S., Liu, C.Y., Chen, T.H., 2008. Adiabatic shearing behavior of different steels under extreme high shear loading. Journal of
Nuclear Materials 374, 313319.
Levitas, V.I., Idesman, A.V., Olson, G.B., 1998. Continuum modeling of strain-induced martensitic transformation at shear-band
intersections. Acta Materialia 47 (1), 219233.
Lichtenfeld, J.A., Mataya, M.C., Van Tyne, C.J., 2006. Effect of strain rate on stressstrain behavior of alloy 309 and 304L austenitic
stainless steel. Metallurgical and Materials Transactions A 37 (1), 147161.
Liu, B., Qiu, X., Huang, Y., Hwang, K.C., Li, M., Liu, C., 2003. The size effect on void growth in ductile materials. Journal of
Mechanics and Physics and Solids 51, 11711187.
Llorca, J., Needleman, A., Suresh, S., 1991. An analysis of the effects of matrix void growth on deformation and ductility in metal
ceramic composites. Acta Metallurgica 39 (10), 23172335.
Low, J.R., Jr., Van Stone, R.H., Merchant, R.H., 1972. NASA Tech. Rep. No. 2, Research Grant NGR 38-087-003, Carnegie-Mellon
Univ.
Lula, R.A., 1966. Stainless Steel. American Society of Materials, 15 pp.
Mangonon, P.L., Thomas, G., 1970. A. The martensite phases in 304 stainless steel. Metallurgical Transactions A 1 (6), 1577
1586.
Maxwell, P.C., Goldberg, A., Shyne, J.C., 1974a. Inuence of martensite formed during deformation on the mechanical behavior of
FeNiC alloys. Metallurgical Transaction A (5), 13191324.
Maxwell, P.C., Goldberg, A., Shyne, J.C., 1974b. Stress assisted and strain-induced martensites in FeNiC alloys. Metallurgical
Transaction A 5, 13051318.
McClintock, F.A., 1968. A criterion for ductile fracture by growth of holes. Journal of Applied Mechanics 35, 363371.
Miura, H., Sakai, T., Okonogi, M., Yoshinaga, N., 2008. Deformation behaviour of carbon steel with dispersed ne voids at
elevated temperatures. Materials Science and Engineering A (483484), 590593.
Monchiet, V., Cazacu, O., Charkaluk, E., Kondo, D., 2008. Macroscopic yield criteria for plastic anisotropic materials containing
spheroidal voids. International Journal of Plasticity 24, 11581189.
Mori, T., Oliver, E.C., Daymond, M.R., Withers, P.J., 2004. Micromechanics of stress-induced martensitic transformation.
Materials Science and Engineering A 378 (12), 479483.
Murr, L.E., Staudhammer, K.P., Hecker, S.S., 1982. Effects of strain state and strain rate on deformation-induced transformation
in 304 stainless steel: Part II. Microstructural study. Metallurgical Transactions A 13 (4), 627635.
Nagy, E., Mertinger, V., Tranta, F., Solyom, J., 2004. Deformation induced martensitic transformation in stainless steels. Materials
Science and Engineering A 378, 308313.

2246

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

Narutani, T., Olson, G.B., Cohen, M., 1982. Constitutive ow relations for austenitic steels during strain-induced martensitic
transformation. Journal de Physique 43 (C4), 429434.
Narutani, T., 1989. Effect of deformation-induced martensitic transformation on the plastic behavior of metastable austenitic
stainless steel. Materials Transactions JIM 30 (1), 3345.
Oberste-Brandenburg, C., Bruhns, O.T., 2004. A tensorial description of the transformation kinetics of the martensitic phase
transformation. International Journal of Plasticity 20, 20832109.
Olson, G.B., Cohen, M., 1972. A mechanism for the strain-induced nucleation of martensitic transformations. Journal of LessCommon Metals 28 (1), 107118.
Olson, G.B., Cohen, M., 1976a. A general mechanism of martensitic nucleation: Part I. General concepts and the FCC ? HCP
transformation. Metallurgical Transactions A 7 (12), 18971904.
Olson, G.B., Cohen, M., 1976b. A General mechanism of martensitic nucleation: Part II. FCC ? BCC and other martensitic
transformations. Metallurgical Transactions A 7 (12), 19051914.
Pardoen, T., Hutchinson, J.W., 2000. An extended model for void growth and coalescence. Journal of Mechanics and Physics of
Solids 48 (12), 24672512.
Pardoen, T., Hutchinson, J.W., 2003. Micromechanics-based model for trends in toughness of ductile materials. Acta Materialia
51, 133148.
Petit, B., Gey, N., Cherkaoui, M., Bolle, B., Humbert, M., 2007. Deformation behavior and microstructure/texture evolution of an
annealed 304 AISI stainless steel sheet. Experimental and micromechanical modeling. International Journal of Plasticity 23,
323341.
Poruks, P., Yakubtsov, I., Boyd, J.D., 2006. Martensiteferrite interface strength in a low-carbon bainitic steel. Scripta Materialia
54, 4145.
Qiu, Y.P., Weng, G.J., 1993. Plastic potential and yield function of porous materials with aligned and randomly oriented
spheroidal voids. International Journal of Plasticity 9, 271290.
Quang, H.L., He, Q.C., 2008. Variational principles and bounds for elastic inhomogeneous materials with coherent imperfect
interfaces. Mechanics of Materials 40, 865884.
Rice, J.R., Tracey, D.M., 1969. On the ductile enlargement of voids in triaxial stress elds. Journal of Mechanics and Physics of
Solids 17, 201217.
Rivas, J.M., Zurek, A.K., Thissell, W.R., Tonks, D.L., Hixson, R.S., 2000. Quantitative description of damage evolution in ductile
fracture of tantalum. Metallurgical Transactions A 31, 845851.
Rizal, S., Homma, H., 2000. Dimple fracture under short pulse loading. International Journal of Impact Engineering 24, 6983.
Salemi, A., Abdollah-zadah, 2008. The effect of tempering temperature on the mechanical properties and fracture morphology of
a NiCrMoV steel. Materials Characterisation 59, 484487.
Sanchez, P.J., Huespe, A.E., Oliver, J., 2008. On some topics for the numerical simulation of ductile fracture. International Journal
of Plasticity 24, 10081038.
Serri, J., Martiny, M., Ferron, G., 2005a. Finite element analysis of the effects of martensitic phase transformation in TRIP steel
sheet forming. International Journal of Mechanical Sciences 47 (6), 884901.
Shrinivas, V., Varma, S.K., Murr, L., 1995. Deformation-induced martensitic characteristics in 304 and 316 stainless steels during
room-temperature rolling. Metallurgical and Materials Transactions A 26 (3), 661671.
Spencer, K., Embury, J.D., Conlon, K.T., Vron, M., Brchet, Y., 2004. Strengthening via the formation of strain-induced martensite
in stainless steels. Materials Science and Engineering A (387389), 873881.
Staudhammer, K.P., Murr, L.E., Hecker, S.S., 1983. Nucleation and evolution of strain-induced martensitic (B.C.C.) embryos and
substructure in stainless steel: a transmission electron microscopy study. Acta Metallurgica 31 (2), 267274.
Stringfellow, R.G., Parks, D.M., Olson, G.B., 1992. A constitutive model for transformation plasticity accompanying straininduced martensitic transformation in metastable austenitic steels. Acta Metallurgica et Materialia 40 (7), 17031716.
Stupkiewicz, S., Petryk, H., 2002. Modelling of laminated microstructures in stress-induced martensitic transformations. Journal
of the Mechanics and Physics of Solids 50 (11), 23032331.
Suzuki, T., Kojima, H., Suzuki, K., Hashimoto, T., Koike, S., Ichihara, M., 1977. An experimental study of the martensite nucleation
and growth in 18/8 stainless steel. Acta Metallurgica 25 (10), 11511162.
Talonen, J., 2007. Effect of strain-induced ((a0 -martensite transformation on mechanical properties of metastable austenitic
stainless steels. Doctoral Dissertation.
Tamura, I., 1982. Deformation-induced martensitic transformation and transformation-induced plasticity in steels. Materials
Science 16 (5), 245253.
Tanaka, K., Oberaigner, E. R., Fischer, F.D., 1994. A unied theory on thermomechanical mesoscopic behaviour of alloy materials
in the process of martensitic transformation. Mechanics of Phase Transformations and Shape Memory Alloys (ASME) AMD
189/PVP 292, pp. 151157.
Tanaka, K., Oberaigner, E.R., Fischer, F.D., 1996. Kinetics on the micro- and macro-levels in polycrystalline alloy materials during
martensitic transformation. Acta Mechanica 116, 171186.
Tang, M., Zhang, J.H., Hsu, T.Y., 2002. One-dimensional model of martensitic transformations. Acta Materialia 50 (3), 467474.
Thomason, P.F., 1990. Ductile Fracture of Metals. Pergamon Press, Oxford.
Tomita, Y., Iwamoto, T., 1995. Constitutive modeling of trip steel and its application to the improvement of mechanical
properties. International Journal of Mechanical Sciences 37, 12951305.
Turteltaub, S., Suiker, A.S.J., 2005. Transformation-induced plasticity in ferrous alloys. Journal of the Mechanics and Physics of
Solids 53, 17471788.
Tvergaard, V., 1981. Inuence of voids on shear band instability under plane strain condition. International Journal of Fracture
17, 389407.
Tvergaard, V., Needleman, A., 1984. Analysis of the cup-cone fracture in a round tensile bar. Acta Metallurgica 32, 157169.
Tvergaard, V., 1990. Material failure by void growth to coalescence. Advance Applied Mechanics 27, 83151.
Venables, J.A., 1962. The martensite transformation in stainless steel. Philosophical Magazine 7 (73), 3544.
Wang, H.S., Yang, J.R., Bhadeshia, H.K.D.H., 2005. Characterisation of severely deformed austenitic stainless steel wire. Materials
Science and Technology 21 (11), 13231328.

A. Das, S. Tarafder / International Journal of Plasticity 25 (2009) 22222247

2247

Wen, J., Huang, Y., Hwang, K.C., Liu, C., Li, M., 2005. The modied Gurson model accounting for the void size effect. International
Journal of Plasticity 21, 381395.
Wu, X.Q., Kim, I.S., 2003. Effects of strain rate and temperature on tensile behavior of hydrogen-charged SA508 Cl.3 pressure
vessel steel. Materials Science and Engineering A 348, 309318.
Xiao, X., Fang, S., Xia, L., li, W., Hua, Q., Dong, Y., 2003. Effect of strain rate on the fracture morphologies of Zr-based bulk metallic
glasses. Journal of Non-Crystalline Solids 330, 242247.
Xue, Q., Cerreta III, E.K., Gray, G.T., 2007. Microstructural characterisation of post-shear localisation in cold rolled 316 L stainless
steel. Acta Materialia 55 (2), 691704.
Zairi, F., Nait-Abdelaziz, M., Gloaguen, J.M., Bouaziz, A., Lefebvre, J.M., 2008. Micromechanical modelling and simulation of
chopped random ber reinforced polymer composites with progressive debonding damage. International Journal of Solids
and Structures 45, 52205236.
Zhao, L., Van Dijk, N.H., Brck, E., Sietsma, J., Van der Zwaag, S., 2001. Magnetic and X-ray diffraction measurements for the
determination of retained austenite in TRIP steels. Materials Science and Engineering A 313 (1), 145152.
Zurek, K., Meyers, M.A., 1970. In: Davison, L., Grady, D.E., Shahinpoor, M. (Eds.), High-pressure Shock Compression of Solids II.
Springer, New York, pp. 2570.

Vous aimerez peut-être aussi