Sie sind auf Seite 1von 10

J Mater Sci (2013) 48:61576166

DOI 10.1007/s10853-013-7412-8

Strain-induced martensite formation in austenitic stainless steel


Mitsuhiro Okayasu Hironobu Fukui
Hiroaki Ohfuji Tetsuro Shiraishi

Received: 28 November 2012 / Accepted: 27 April 2013 / Published online: 8 May 2013
Springer Science+Business Media New York 2013

Abstract In situ measurements of the strain-induced


martensitic transformation (SMTs) of SUS304 stainless
steel that takes place during tensile loading at room temperature were performed around the notch of a dumbbellshaped specimen where high stress concentration occurs.
Even in the low plastic strain regime, with loading to 0.2 %
proof stress (r0.2), some SMTs occurred. However, the area
fraction of the Fe-a0 -martensite phase did not increase
significantly even when the sample was loaded to the
ultimate tensile strength (rUTS). After the rUTS point, the
total fraction of the Fe-a0 phase increased dramatically to
the fracture point (rf). The phase textures of Fe-a0 and Fe-c
were almost equal at (rUTS - rf)/2, and the Fe-a0 phase
was observed over almost the entire measurement area
around the notch at the rf point. However, the area fraction
of the Fe-a0 phase at the rf point decreased far away from
the fracture surface, to an extent that the total fraction of
the Fe-a0 phase was almost the same as that of the Fe-c
phase in an area about 1.7 mm from the fracture face.
Different martensite characteristics were detected in the
stainless steel, depending on the applied load level. This
was attributed to the severity of deformation. In particular,
deformation twinning, created around rUTS, and severe
plastic deformation before fracture make a strong Fe-a0
phase. Details of this phenomenon are interpreted using
various approaches, including electron backscatter diffraction analysis and finite element analysis.
M. Okayasu (&)  H. Fukui  T. Shiraishi
Department of Materials Science and Engineering, Ehime
University, 3 Bunkyo-cho, Matsuyama, Ehime 790-8577, Japan
e-mail: okayasu.mitsuhiro.mj@ehime-u.ac.jp
H. Ohfuji
Geodynamics Research Center, Ehime University,
2-5 Bunkyo-cho, Matsuyama, Ehime 790-8577, Japan

Introduction
Austenitic stainless steels have received special attention
because of their use for various engineering components,
such as in power plants and the automobile industries, owing
to their excellent machinability, high corrosion resistance,
high strength and high ductility. However, although
austenitic stainless steels have high material ductility, this is
significantly affected by strain-induced martensitic transformations (SMTs). A variety of stainless steel components
are produced by mechanical processes including stretching,
drawing and bending, all of which cause severe plastic
deformation [1]. Experimental and numerical approaches
have been used to attempt to understand the stressstrain
characteristics in stainless steel components [2, 3]. It is
believed that austenite phases in some stainless steels are
metastable owing to the presence and amount of Cr and Ni,
both of which lead to a SMT when the stainless steel is
subjected to plastic deformation. The life of stainless steel
components in service is generally a function of the severity
of this plastic deformation. Consequently, an examination of
plastic strain characteristics in these stainless steels is of
considerable importance, and several techniques have been
proposed to observe the localised plastic deformation zone
[4]. It is believed that austenitic stainless steels exhibit significant work hardening, resulting in the transformation from
metastable austenite to martensite [5]. It is also considered
that a marked increase in elongation occurs when martensites
are formed during the deformation, which is called transformation-induced plasticity [6]. The volume fraction of
austenite to martensite transformation depends on the strain
level, temperature and strain rate [3].
Zong-yu et al. [7] investigated the influence of pretransformed martensite on the work hardening behaviour of
metastable SUS304 austenitic stainless steel. One of their

123

6158

results was that the work hardening exponent of the


stainless steel decreases with an increase of pre-transformed martensite. SUS304 stainless steel is a typical
austenitic stainless steel and is known to exhibit considerable
work hardening; hence a pre-strained-type stainless steel
would be expected to show a martensitic transformation
[1]. In the study by Dan et al. [5], a constitutive model for the
transformation-induced plasticity accompanying the SMT
was developed. The increase in the number of nucleation
sites in austenite owing to plastic deformation is formulated
as an increase of the shear band intersection. The nucleation
site probability is dependent not only on the stress state,
plasticity strain and temperature of the environment but also
on the strain rate. During their analysis of the tensile process
at a strain rate of 0.001 s-1, an increase in temperature led to
a slow martensite transformation. The shear band, driving
force and parameter-a0 transformation also decrease when
the deformation temperature is higher [5]. This is because the
shear band intersection is reduced at higher temperature,
resulting in weak plasticity-induced transformation. Beese
and Mohr examined the effect of the stress state on the
transformation kinetics of stainless steel sheets (301LN) at
room temperature using newly developed experimental
techniques for simple shear and large strain in-plane compression. Their experimental results indicated that the martensite transformation kinetics cannot be described solely by
a monotonically increasing function of stress triaxiality. For
instance, less martensite is developed under equibiaxial
tension compared with uniaxial tension for the same increase
in equivalent plastic strain [8].
To date, several studies have systematically investigated
the SMT characteristics of austenitic stainless steels, along
with relevant information used to explain their material
characteristics. Although some significant information has
been reported, it is believed that further study is required to
obtain more effective SMT characteristics for these stainless
steels. This is especially true, as the SMT characteristics
have been examined mainly using X-ray diffraction, TEM
and magnetic permeability measurements, [911] which
may not show clear phase textures, e.g. with two-dimensional images. In the present study, in situ measurements of
SMT behaviour were made during tensile loading, and the
effects of deformation and crystal orientation characteristics
on the SMT behaviour were investigated using electron
backscatter diffraction (EBSD) analysis.

J Mater Sci (2013) 48:61576166

chemical
composition
FeNi8.6Cr18.4Si0.45Mn1.8
Mo0.3S0.03C0.08 (wt%). The SUS304 was annealed at
1150 C for 2 h and air-cooled to remove any residual
stresses. The stainless steel consisted of an austenite phase
with a mean grain size of 92 lm in diameter. Figure 1
shows a test specimen. The specimens were machined into
the form of rectangular dumbbell shapes with two notches.
The reason for the notches was to measure the SMT
characteristics in the same area and similar stress level.
Specimens were machined by a wire electric discharge
machine. The stress concentration rate in the specimens
around the notches was about Kt = 2.4 [12]. In this case,
the specimen surfaces were finished with a smooth level of
Ra = 1.6 to reduce the stress concentration arising from
tool marks. It should be pointed out that the material
properties of the notched specimens are different from
those of standard smooth tensile specimens. In this case,
the ductility of the sample would be reduced by the notch
effect [13].
Tensile tests were conducted using an electro-servohydraulic system with 50 kN capacity. The applied load
and strain were measured with a commercial load cell and
a strain gauge, respectively. After a specimen was loaded
to several specific loading points, the SMT characteristics
of the specimen were examined adjacent to the notch tip
(A0 , see Fig. 1), where relatively high stress concentration
occurs. Note first that the area A0 does not have the highest
stress concentration, owing to the area being about 1 mm
away from the notch tip. The reason for this measurement
was to obtain clear SMT characteristics during the tensile
tests. In fact a clear SMT cannot be obtained in the area of
the sample adjacent to the notch tip (the highest stress
concentration area) because of severe material damage.
This examination was carried out only at room temperature
(about 293 K) as a first approach even though the SMT is
sensitive to sample temperature [14]. The crystal
175
57

6.5

12.5

R12.5

Experimental procedures

2
R1
1

123

t=5

Materials and experiments


The study reported used a commercial SUS304 stainless
steel plate, produced by a hot-rolling process, with

20

Measurement area (A)

Fig. 1 Schematic illustration of the test specimen

J Mater Sci (2013) 48:61576166

Finite element analysis

700

E
600

Tensile stress, MPa

orientation characteristics were investigated by EBSD


analysis, using a high-resolution electron microscope
(JSM-7000F, JEOL Ltd). Based upon an analysis of the
EBSD patterns, the phase textures (Fe-a0 and Fe-c phases)
of the matrix were determined using HKL Channel 5
software. The main measurement conditions for this analysis were as follows: the sample was tilted to an angle of
70, and an accelerating voltage of 15 kV, beam current of
5 nA and measurement step size of 4 lm were used. The
sample surfaces for this examination were polished for
about 2 h to mirror level in a vibro-polisher using colloidal
silica. The polishing process was executed carefully with a
low applied load so as to reduce sample damage.

6159

400

300

200

100

0
0

To investigate the stressstrain distribution in the test


specimen of SUS304 stainless steel after tensile loading,
finite element analysis was conducted using ANSYS software. In this analysis, a two-dimensional finite element
simulation with 8-node quad elements was employed. The
finite element analysis model was designed to simulate the
actual specimen, as shown in Fig. 1. The mesh size of the
model was set to 0.05 mm, giving a total of 3251 quadrilateral solid elements. To identify the plastic strain levels in
this analysis, the option of bilinear kinematic hardening
was selected, in which the initial slope of the stressstrain
relationship is taken as the Youngs modulus of the
material. After the yield point, the curve continues along a
second slope defined by the tangent to the curve. Based on
our experimental results, the following material properties
were determined: Youngs modulus E = 210 GPa, tangent
modulus T = 1105 MPa, Poissons ratio m = 0.3 and yield
strength r0.2 = 290 MPa. Note that, in this case, the yield
strength was defined as the stress value at which 0.2 %
permanent deformation occurs.

Results and discussion


Figure 2 displays representative tensile stress (r)strain (e)
curves for our SUS304 stainless steel. Because the specimen has notches, its tensile properties are different from
those of a standard smooth specimen, i.e. there is a notch
effect. We have examined the tensile properties of the
smooth specimen in addition to those of the notched one,
and it appears that both ultimate tensile strength and
Youngs modulus are almost the same, but the strain to
failure for the notched specimen is slightly lower compared
to the smooth specimen. This is because the presence of
triaxial stress field and steep stress (such as arise at a notch)
gradients with the notch effect results in a change of
ductility.

500

10

20

30

40

50

Strain, %
Fig. 2 Tensile stressstrain curve for the SUS304 stainless steel

In the present study, several specimens were loaded to


the specific loading (or strain) points labelled A, B, C, D, E
and F in Fig. 2. Point A is the specimen without any
applied load, i.e. 0 N. Points B, D and F are the yield point
r0.2, ultimate tensile strength rUTS and fracture point rf,
respectively. Finally, points C and E are at load levels of
(rUTS - r0.2)/2 and (rUTS - rf)/2, respectively. It should
be noted that, in the tensile test, the applied load was
applied at a low strain rate of about 0.0005 s-1, which is
much lower than that used in previous studies [15, 16]. The

Sample A

Sample B

Sample C

Sample D

Sample E

Sample F
10mm

Fig. 3 Photographs of the test specimens after loading to the specific


points shown in Fig. 2

123

6160

J Mater Sci (2013) 48:61576166

reasons behind the low strain rate are (i) to create a clear
SMT [3, 17] and (ii) to easily control the load level at the
specific points (points AF). The effect of strain rate on
SMT characteristics has been investigated by Das et al.
[17]. One of their conclusions is that the total volume
fraction of martensite decreases when the strain rate is
Sample A

increased because of the enhancement of sample temperature. Talonen et al. [18] have also similarly reported that
more strain-induced a0 -martensite is formed in steel at a
lower strain rate and higher Md30 temperature. The Md30
temperature of our stainless steel approximated using
Gladmans formula is 11.7 C, i.e. 497462(C ? N)
Sample B

Sample C

Sample D

Sample E

Sample F

111

50m

001

101

Fig. 4 Crystal orientation maps of the SUS304 stainless steels after loading to the specific points shown in Fig. 2 (Color figure online)

123

J Mater Sci (2013) 48:61576166

6161

9.2(Si)8.1(Mn)13.7(Cr)20(Ni)18.5(Mo), which is
similar to that for a related stainless steel [19]. However, an
opposite trend for SMT is obtained in the stainless steel
after impact loading tests, in which the volume fraction of
martensite increases with increasing strain rate, i.e.
8 9 102 to 4.8 9 103 s-1 [20]. The increase in martensite
volume fraction caused by the increase of strain rate can
lead to high strength, where a high dislocation density

would be introduced into the austenite to accommodate the


martensite volume expansion [21]. Figure 3 shows photographs of the specimens after loading to the specific points
mentioned above. The different degrees of permanent
deformation around the notches can clearly be seen,
especially after the material yields (samples CF), while no
clear deformation can be observed in sample B. In sample
D, severe deformation occurs because it undergoes necking

Sample A

Sample B

Sample C

Sample D

Sample E

Sample F

-phase
50m

-phase

Fig. 5 Orientation imaging microscopy maps for samples AF (Color figure online)

123

6162

123

used in previous studies to alter the texture are close to our


result, further study will be required, because notched
specimens were employed in our approach. A different result
was also reported by Spencer et al. [23], where the stainless
steel was virtually saturated with a0 -martensite before
necking (i.e. r \ ultimate tensile strength), which may be
affected by the experiments being conducted at the lower
temperature of 77 K [12, 23]. An interesting approach was
used by Hecker et al. [24]. They investigated the SMT
characteristics for 304 stainless steel using the von Mises
effective strain criterion, which gives a reasonable correlation of transformation kinetics under general strain states: the
principal effect of increased strain rate was obtained at
strains greater than 0.25 [24].
Consideration should be given as to whether microcracks affect SMT or not in the present study. In general,
the micro-cracks would be created by the following process: micro-voids are created during the earlier failure
process, and their coalescence occurs as the load becomes
greater than the ultimate tensile strength [25]. Because of
stress relaxation, caused by micro-crack generation, this
could be a significant factor in the SMT characteristics. It is
considered that micro-cracks are created during the tensile
tests especially in samples E and F. Since the data collection rate in the EBSD analysis was decreased for the
samples loaded to points E and F as mentioned for Figs. 4
and 5, this occurrence may be attributed to the microcracks. However, further study will be required in the future.
It is also considered that the nucleation of a0 -martensite
is always confined to microscopic shear bands, including
strain, faults and twin intersections [11]. Details of those
microstructural effects on the martensite transformation are
discussed below.
1.0

Rate of area fraction for - and -phase

[12], and more severe deformation occurs in samples E and


F. In previous reports, similar loadstrain relationships
were obtained for the appropriate stainless steel, where
high strain increases the severity of the martensitic transformation [22].
Figure 4 depicts the crystal orientation maps obtained at
points AF. The colour level of each pixel in the crystal
orientation map is defined according to the deviation of the
orientation measured from the direction parallel to ND (see
the colour key of the stereographic projection). Moreover,
the dark areas in this analysis are associated with regions in
which data cannot be collected. This is because of the
severe material damage, e.g. collapse of microstructure and
micro-cracks. Austenite grains can be clearly seen in the
SUS304 stainless steel when it is loaded to an extent less
than the ultimate tensile strength. On the other hand, some
twinning-like deformation zones are observed in sample D,
as indicated by the circles. The grain formations in samples
E and F are seen to have collapsed to tiny regions and
randomly distributed crystal orientation caused by more
severe slip deformation. Basically, the crystal orientation in
samples AD is different depending on the grain, whereas
this cannot be seen clearly in samples E and F.
Figure 5 shows orientation imaging microscopy maps for
samples AF, and Fig. 6 shows the variations of the area
fraction of Fe-a0 -martensite and Fe-c phases as a function of
tensile strain. Note that the measurement areas displayed in
Fig. 5 are the same as those shown in Fig. 4. It is clear that a
strong texture of Fe-c is apparent until the samples are
strained to about 40 %, at which point the total area fraction
of Fe-c is about 85 % (samples AD). It is interesting to
mention first that even if a relatively high strain (e = 40 %)
is present in sample D, the SMT is weak. In addition, the
characteristics of Fe-a0 -martensite formation for sample D
(Fig. 5) are different from those for samples A, B and C. For
example, several Fe-a0 phase colonies are obtained for
sample D, and the Fe-a0 phase can be detected, with many
tiny regions (dot-like formations), for samples B and C, as
indicated by the dashed circles. This contrasting martensitic
characteristic will be discussed later. In the present study, the
total area fraction of the Fe-a0 phase apparently increases
with increasing strain level beyond e = 40 % (or rUTS),
during which the phase textures of Fe-a0 and Fe-c are similar
(Fe-a0 = 55 % and Fe-c = 45 %) at e = 45 % (sample E).
In sample F, almost the whole measurement area is characterised by the presence of the Fe-a0 -martensite phase
(Fe-a0 & 80 %). Interestingly, corresponding results were
reported by Ogata et al. [14], where there is no clear SMT in
SUS316L until a strain level of 40 %. In another report, the
volume fraction of martensite increased sharply when the
sample was strained to about 2550 %, which was dependent
on the strain rate: the higher the strain rate, the lower the
volume fraction of martensite [17]. Although the strain levels

J Mater Sci (2013) 48:61576166

0.9

C
F

0.8
0.7

-phase

0.6

-phase

0.5

0.4
0.3

0.2
0.1
0

A
0

10

20

30

40

50

Strain, %
Fig. 6 Variation of the area fraction of Fe-a0 -martensite and Fe-c
phase as a function of the tensile strain, obtained on the basis of Fig. 5

J Mater Sci (2013) 48:61576166

6163

Sample A

Sample B

Sample C

Sample D

Sample E

Sample F

7
50m

Fig. 7 Strain distribution defined with crystal orientation angle (Color figure online)

Figure 7 depicts the lattice strain distribution in samples


AF. In this case, the lattice strain was defined by the
crystal orientation angle. It is seen that there are different
strain distribution characteristics. Compared to the orientation imaging microscopy maps in Fig. 5, the severe strain
is randomly distributed in the area of Fe-a0 phase, which
may be associated with severe plastic deformation [6].
In addition, lattice strain, which is related to SMT, is

enhanced in twins, although that strain is at a low level as


indicated by the dashed circles. From this result, SMT can
be attributed to two different strain patterns: plastic
deformation and twin formation. Note that although there
are deformation twins in samples B and C as indicated by
the circles, those are not attributed to SMT [26]. This is
because those twins are not created during the tensile
loading.

123

6164

J Mater Sci (2013) 48:61576166


Measurement zone

1mm
Crack

Crack surface

-phase

0.7 mm

0.1mm

-phase

Fig. 8 Orientation imaging microscopy map in the SUS304 stainless steels in the region of the crack surface (Color figure online)

123

room temperature [27], where the formation of the


mechanical twin structure was detected at a strain level of
e = 0.4 [27]. This result is similar to that for sample D (at
rUTS). Note that a Luders band front may have given significant SMT because of the high initial dislocation density
especially in samples D and E [23], but this cannot be
clarified in the present study, and will be studied in the
future.
The Fe-a0 and Fe-c phases were further investigated
after the tensile tests. Figure 8 shows the orientation
imaging microscopy map recorded for sample F around the
notch and fracture surface (crack), and Fig. 9 shows the
variation of the area fraction of the Fe-a0 and Fe-c phases
as a function of the distance from the fracture surface. A
high concentration of the Fe-a0 -martensite phase is detected near the crack, which could be a result of the high
plastic strain. The area fraction of the Fe-a0 phase
1.0

Rate of area fraction for - and -phase

Several researchers have investigated the associated


SMT characteristics of SUS304 stainless steels. In the
study by Shen et al. [15] it is made clear that a high fraction
of martensite phase is obtained in related stainless steels
when loaded at a lower strain rate of 3 9 10-3. Furthermore, they found, using careful TEM observations, that
stacking faults and twins preferentially occur before martensitic nucleation and both e-martensite and a0 -martensite
are observed in the deformed microstructures. The presence of a0 -martensite is attributed to the twinning deformation and the presence of e-martensite to the plastic
deformation [15]. It is generally considered that the twinning deformation occurs most readily in steels subjected to
high strain rate and/or low temperature conditions [12].
The observation of SMT, obtained with the low strain rate
of 3 9 10-3 in the study by Shen et al. [15], might be
related to the texture of the e-martensite. From the above
information, it may be considered that strong SMT,
obtained in previous study after tensile tests at a lower
temperature [22], is influenced by the twinning deformation, thus resulting in a0 -martensite formation [15, 22].
Interesting results were further reported by Shen et al.
[15], who showed that different kinds of martensite can be
formed. Alpha prime martensite is mainly formed when
high loads are applied to the sample, e.g. loads greater than
the ultimate tensile strength, as for our samples D, E and F.
In contrast, e-martensite is observed when the loading is
less than about (rUTS - r0.2)/2, which is the case for our
samples B and C. As mentioned in the discussion of Fig. 5,
the texture of Fe-a0 -martensite formation is different in our
SUS304 when it is loaded with more than or less than the
ultimate tensile strength (samples B and C vs. sample D).
This difference in texture may be attributed to the different
degree of martensite formation. In addition, for sample D,
the Fe-a0 -martensite phases seem to be distributed mainly
in the twinning deformation zones, as evident from Figs. 4
and 5. It has been reported that mechanical twinning could
give a0 -martensite resulting from uniaxial tensile tests at

0.9
0.8
0.7

-phase

0.6

-phase

0.5
0.4
0.3
0.2
0.1
0

Distance from the crack surface, mm


Fig. 9 Variation of the area fraction of Fe-a0 and Fe-c phase as a
function of the distance from the crack surface, obtained on the basis
of Fig. 8

J Mater Sci (2013) 48:61576166

6165

Fig. 10 Plastic strain (x-axis)


distribution obtained by FE
analysis (Color figure online)

A
1 mm

x-axis

0.05

0.3

0.5

(1)

0.75

1.2

1.4

1.7

2.0

Even if low strain at the r0.2 point occurs, a slight


increase of the Fe-a0 -martensite fraction is obtained.
The total area fraction of the Fe-a0 -martensite phase
at the rUTS point is almost the same as that at the r0.2
point. After loading at greater than rUTS (or
e [ 40 %), the area fraction of the Fe-a0 phase
increases dramatically, although the phase textures of
Fe-a0 and Fe-c are almost the same as at (rUTS - rf)/2.
In addition, the Fe-a0 -martensite phase is formed over
almost the entire region at rf.
The characteristics of the Fe-a0 -martensite phase are
apparently different, depending on the sample. For
example, many tiny fractions of the Fe-a0 phase are
distributed in the un-twinning deformation zones
when the sample is subjected to a load of less than
(rUTS - r0.2)/2, whereas several Fe-a0 -martensite
colonies are formed in the twinning deformation
zones when the load is greater than rUTS. In this case,
two main factors are observed to affect the Fe-a0 martensite phase: (i) twinning deformation and (ii)
severe plastic deformation. In particular, deformation
twinning is created around rUTS and severe plastic
deformation is detected before fracture.
A high area fraction of the Fe-a0 -martensite phase is
detected near the notch, but the area fraction of the
Fe-a0 phase decreases in the area far away from the
crack surface. The total fraction of the Fe-a0 phase is
almost the same as that of the Fe-c phase in an area
about 1.7 mm from the crack surface.

decreases in the area far away from the crack face. The
total area fraction of the Fe-a0 phase is almost the same as
that of the Fe-c phase in the area about 1.7 mm from the
crack face. In the study by Huang et al. [21] variation of the
area fraction of martensite was similarly examined, and
the area fraction of Fe-a0 and Fe-c phases was almost the
same in the area about 5 mm away from the fracture surface,
which is slightly different from our findings. This may be
attributed to the different specimen geometry, namely with
or without notches. An associated experiment was also
carried out by Nakajima et al. They investigated the
occurrence of SMT in SUS304 stainless steels after fatigue
tests (cyclic loading), but no clear SMT was detected in their
sample adjacent to fatigue cracks [28]. The reason behind
this may be the weaker plastic deformation, arising from the
high cyclic loading speed of 53 Hz.
Figure 10 displays the plastic strain distribution
obtained by finite element analysis. In this analysis, the
specimen was loaded to its ultimate tensile strength of
695.9 MPa, i.e. rUTS. This analysis reveals high plastic
strain in the sample adjacent to the notch owing to the high
stress concentration, and the plastic strain level decreases
in the area far away from the notch (Fig. 10). The value of
the plastic strain in area A0 (Fig. 10) is found to be about
48 %, and this value is relatively close to the experimentally obtained strain level after loading with more than the
ultimate tensile strength, e.g. points E and F (Fig. 2). Note
that area A0 is the same area as that used for the EBSD
measurements (Fig. 1). This numerical approach suggests
that SMT in associated SUS304 stainless steels can occur
to a great degree when the sample is strained by more than
40 %.

Acknowledgement This study was technically supported by Mr.


Yuki Sato at Akita Prefectural University in Japan.

Conclusions

References

A SMTs in SUS304 stainless steel near a notch was


examined using an electron backscattering diffraction
approach. The results obtained were as follows:

(2)

(3)

1. Nakajima M, Akita M, Uematsu Y, Tokaji K (2010) Proc Eng


2:323
2. Jia N, Peng RL, Chai GC, Johansson S, Wang YD (2008) Mater
Sci Eng A 491:425

123

6166
3. Chen X, Wang Y, Gong M, Xia Y (2004) J Mater Sci 39:4869.
doi:10.1023/B:JMSC.0000035327.55210.99
4. Okayasu M, Sato K, Takasu S (2010) J Mater Sci 45:1220. doi:
10.1007/s10853-009-4068-5
5. Dan WJ, Zhang WG, Li SH, Lin ZQ (2007) Comput Mater Sci
40:101
6. Tamura I (1982) Met Sci 16:245
7. Zong-yu X, Sheng Z, Xi-Cheng W (2010) J Iron Steel Res Int
17:51
8. Beese AM, Mohr D (2011) Acta Mater 59:2589
9. Zhang HW, Hei ZK, Liu G, Lu J, Lu K (2003) Acta Mater
51:1871
10. Varma SK, Kalyanam J, Murr LE, Srinivas V (1994) J Mater Sci
Lett 13:107
11. Murr LE, Staudhammer KP, Hecker SS (1982) Metall Trans A
13A:627
12. Hertzberg RW (1996) Deformation and fracture mechanics of
engineering materials, 4th edn. Wiley, New York, p 18
13. Dieter GE (1986) Mechanical metallurgy, 3rd edn. McGraw-Hill,
Inc, New York, p 314
14. Ogata T, Yuri T, Ono Y, Cryo J (2007) Soc Jpn 42:10 in Japanese
15. Shen YF, Li XX, Sun X, Wang YD, Zuo L (2012) Mater Sci Eng
A 552:514

123

J Mater Sci (2013) 48:61576166


16. Mirzadeh H, Najafizadeh A (2010) Mater Sci Eng A 527:1856
17. Das A, Sivaprasad S, Ghosh M, Chakraborti PC, Tarafder S
(2008) Mater Sci Eng A 486:283
18. Talonen J, Nenonen P, Pape G, Hanninen H (2005) Metall Mater
Trans A 36A:421
19. Nebel Th, Elfler D (2003) Sadhana 28:187
20. Lee W-S, Lin C-F (2000) Scr Mater 43:777
21. Huang GL, Matlock DK, Krauss G (1989) Metall Trans A
20A:1239
22. Muller-Bollenhagen C, Zimmermann M, Christ H-J (2010) Int J
Fatigue 32:936
23. Spencer K, Embury JD, Conlon KT, Veron M, Brechet Y (2004)
Mater Sci Eng A 387389:873
24. Hecker SS, Stout MG, Staudhammer KP, Smith JL (1982) Metall
Trans A 13A:619
25. Benzerga AA, Leblond J-B (2010) Adv Appl Mech 44:169
26. Kinoshita Y, Yardley VA, Tsurekawa S (2011) J Mater Sci
46:4261. doi:10.1007/s10853-010-5241-6
27. Choi J-Y, Jin W (1997) Scr Mater 36:99
28. Nakajima M, Uematsu Y, Kakiuchi T, Akita M, Tokaji K (2011)
Proc Eng 10:299

Das könnte Ihnen auch gefallen