Plastic Deformation and Creep Damage Evaluations of Type 316 Austenitic Stainless Steels by EBSD

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

M A TE RI A L S CH A RACT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3 –9 2 2

available at www.sciencedirect.com

www.elsevier.com/locate/matchar

Plastic deformation and creep damage evaluations of type 316


austenitic stainless steels by EBSD

Rika Yoda a,b,d,⁎, Toshinori Yokomaku c , Nobuhiro Tsuji d


a
Electronics Division, Kobelco Research Institute Inc., 1-5-5 Takatsuka-dai, Nishi-ku, Kobe, Hyogo 651-2271, Japan
b
Department of Adaptive Machine Systems, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita,
Osaka 565-0871, Japan
c
Engineering Mechanics Division, Kobelco Research Institute Inc., 1-5-5 Takatsuka-dai, Nishi-ku, Kobe, Hyogo 651-2271, Japan
d
Department of Materials Science and Engineering, Graduate School of Engineering, Kyoto University, Yoshida Honmachi, Sakyo-ku, Kyoto,
606-8501, Japan

AR TIC LE D ATA ABSTR ACT

Article history: The inspection method of plastic and/or creep deformations has been required as the
Received 10 May 2009 quantitative damage estimation procedure for structural components especially used in
Received in revised form 6 May 2010 electric power plants. In this study, the method using electron backscatter diffraction (EBSD)
Accepted 10 May 2010 was applied to the deformation and damage evaluation of austenitic stainless steels
strained by tension or compression at room temperature and also tested in creep at high
temperature. It was found that the value of Grain Average Misorientation (GAM) which
Keywords: showed the average misorientation for the whole observed area including over several
Electron backscatter diffraction dozen grains, was a very useful parameter for quantifying the microstructural change as
(EBSD) either the plastic or creep strain increased. The unique linear correlation was obtained
Average misorientation between GAM and plastic strain in tension and compression. For creep damage evaluation,
Plastic deformation the difference of grain average misorientation from the value of the unstrained specimen
Creep damage (ΔGAM) showed an excellent correlation with the inelastic strain below strain at which the
Inelastic strain tertiary creep began.
Austenitic stainless steel © 2010 Elsevier Inc. All rights reserved.

1. Introduction tion. Therefore, the inspection method of plastic and/or creep


deformation has been required as the quantitative damage
In structural components for electric power or petro-chemical estimation procedure for above structural components.
plants, residual stresses, caused by plastic deformation in Various methods have been developed for this purpose [1],
metal forming and welding processes, may raise the stress such as hardness measurement, A-parameter method using
corrosion cracking sensitivity or deteriorate the fatigue replicas of surface microstructure, ultrasonic, electric, mag-
properties. Also, thermal and mechanical stresses at service netic, and X-ray methods. Hardness measurement is a very
temperatures may cause the life reduction by creep deforma- simple method, but it cannot distinguish hardening by plastic
deformation from that by precipitation occurred at elevated
temperatures. Surface replication method, such as A-parame-
ter method to evaluate the void fraction, can be applied only to
the materials including creep voids. Either ultrasonic, electric,
⁎ Corresponding author. Electronics Division, Kobelco Research
magnetic or X-ray method, utilizes indirect parameters, such
Institute Inc., 1-5-5 Takatsuka-dai, Nishi-ku, Kobe, Hyogo
651-2271, Japan. Tel.: +81 78 992 6043; fax: +81 78 990 3062. as ultrasonic amplitude, electric resistance, Barkhausen noise,
E-mail addresses: yoda.rika@kki.kobelco.com (R. Yoda), and half breadth value of X-ray diffraction peak, each of which
yokomaku.toshinori@kki.kobelco.com (T. Yokomaku), does not necessarily have a clear physical meaning of the
nobuhiro.tsuji@ky5.ecs.kyoto-u.ac.jp (N. Tsuji). deformation or damage itself.

1044-5803/$ – see front matter © 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.matchar.2010.05.006
914 MA TE RI A L S CH A R A CT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3–9 2 2

On the other hand, a method using electron backscatter Table 2 – Tensile properties of the starting materials.
diffraction (EBSD) detected in a scanning electron microscope Material Yield Tensile Elongation Reduction
(SEM) has the advantage that the degree of deformation or stress strength (%) in area (%)
damage can be expressed quantitatively as a local change in (MPa) (MPa)
crystal orientation of grains[2–13]. Also, the deterioration in
Type 316 287 607 60 74
quality of EBSD patterns is considered to correspond to a change Type 316NG 282 577 53 80
in dislocation density caused by plastic strain [14–18]. The recent
improvement of the equipment and analysis software is propel-
ling the studies regarding damage evaluation by the EBSD
method. It has been reported that local misorientation in grain plastic strain, in order to apply the EBSD method to various
increased with increasing plastic strain [5,19–28]. On the other damage evaluations in real structural materials. Uniaxial
hand, for creep damage evaluation of heat-resistant material, deformation is the most fundamental deformation mode.
Takaku et al. [29] showed that the same tendency was obtained Therefore, both tensile and compression deformations were
even during increasing creep strain in a Ni base superalloy. adopted as uniaxial plastic deformation in this study.
However, Fujiyama et al. [30] and Ohtani et al. [31] clarified that Round-bar tensile specimens 7 mm in diameter and 15 mm
local misorientation decreased with creep strain in the material in gauge length, and cylindrical compression specimens 10 mm
having martensitic structure, such as 10%Cr steel and type 403 in diameter and 20 mm in length were machined from the type
stainless steel, respectively. Furthermore, Mitsuhara et al. [32] 316NG stainless steel plate. These specimens were deformed
reported that three different parameters were suitable for three uniaxially in tension or compression, respectively, at room
different creep regions in a 9–10%Cr steel. These results suggest temperature. Both tensile and compression deformations were
that appropriate EBSD parameters for creep damage evaluation carried out at an initial strain rate of 3.3 × 10− 3 s− 1 using
depend on the kind of materials or phases. SHIMADZU Autograph. The specimens were deformed up to
In this study, first, we evaluate the relationship between four different strain levels: 0.87% engineering strain (true strain:
the Grain Average Misorientation (GAM), that is one of the εpl = 0.0087), 2.82% (εpl = 0.0278), 4.79% (εpl = 0.0468) and 9.70%
parameters obtained by EBSD analysis, and the plastic strain (εpl = 0.0926) for tensile deformation, and 0.88% (εpl = 0.0088),
in the most fundamental tensile and compressive deforma- 2.84% (εpl = 0.0288), 4.86% (εpl = 0.0498) and 9.72% (εpl = 0.102) for
tion, using an austenitic stainless steel. Secondly, we discuss compression, respectively.
the applicability of this parameter as a measure of the creep
damage and deformation. 2.3. Creep Deformation

The round-bar creep specimens of 10 mm in diameter and


2. Experimental Procedures 50 mm in gauge length, were machined from the type 316
stainless steel plate. Creep rupture tests were conducted at
2.1. Materials 600 °C under the applied stresses of 250, 270, 280 and 300 MPa
and the corresponding rupture times (tr) were 2600, 599, 409
Two types of austenitic stainless steels are used in this study. and 156 h, respectively, as is shown in Fig. 2(a). Creep
One is a type 316 stainless steel (JIS-SUS316) that is a common interruption tests were carried out under an applied stress of
material for high-temperature applications. The other is a type 270 MPa (rupture time = 599 h). Interruption times of the tests
316 nuclear grade stainless steel (ASTM type 316NG) for nuclear are 72, 144, 210, 300 and 419 h which correspond to the creep
power plant component. The steels were solution-treated after damage ratio (Dc = t/tr, t; interruption time) of 0.12, 0.24, 0.35,
hot rolling, and provided for the present experiments. The 0.50 and 0.70, respectively. Creep curves of the interrupted
chemical compositions and tensile properties of the steels are specimens are shown in Fig. 2(b). The 316 steel showed typical
summarized in Tables 1 and 2, respectively. The optical creep behaviors.
microstructures of the starting specimens solution-treated are
shown in Fig. 1. Both steels show fully recrystallized micro- 2.4. EBSD Analysis
structures, and the average grain diameters are 106 μm and
23 μm for type 316NG and type 316, respectively. OIM (Orientation Imaging Microscopy™) system of EDAX/TSL
attached to a field-emission SEM (FEI XL30S-FEG) was used for
2.2. Tensile and Compression Deformation the EBSD analysis. The cross-section parallel to the stress axis
in each specimen was polished using progressively finer
The first purpose of this study is to clarify the relationship grades of diamond paste, ranging from 6 µm to 1 µm and
between some parameters obtained by EBSD analysis and finally 0.25 µm in particle size, and then they were lightly

Table 1 – Chemical compositions of the stainless steels studied.


Material (mass%)

C Si Mn P S Cu Ni Cr Mo N

Type 316 0.05 0.26 1.28 0.033 0.029 – 10.0 16.99 2.01
Type 316NG 0.016 0.43 1.54 0.02 0.0006 0.26 11.94 17.13 2.15 0.1
M A TE RI A L S CH A RACT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3 –9 2 2 915

Fig. 2 – (a) Relationship between applied stress and rupture


time in creep tests of type 316 stainless steel at 600 °C. The
data from the interrupted tests are indicated as the symbol
Fig. 1 – Optical microstructures of the (a) type 316NG and
“×”, (b) Creep curves of the type 316 stainless steel specimens
(b) type 316 starting specimens.
tested at 600 °C for various periods under an applied stress of
270 MPa.

etched using colloidal silica to finish it to a smooth and flat


surface. The center area of the polished plane was analyzed.
The same area was also observed by SEM. The EBSD boundary within it. Each pixel i, having a crystal orientation
measurements were done in the area of 1000 × 1000 μm2 with datum, has six neighboring pixels because the orientation
a step of 2 μm on a hexagonal grid for the tensile and mapping was done on the hexagonal grid in this study.
compression test specimens and 240 × 240 μm2 with a step of However, when a neighboring pixel belongs to a different
0.8 μm for the creep test specimens. These conditions were grain, it should be excluded from the calculation of misorien-
selected in order to include a sufficient number of both crystal tation. Therefore, up to six different misorientations exist
grains in the measured region and pixels in the grain. That is, between the pixel i and neighboring pixels within an identical
each measurement area had about 100 grains, and an average grain surrounded by the grain boundary. The number of
number of pixels per grain were about 1000. misorientation data in one grain, m, is defined by the equation,
The Grain Average Misorientation (GAM), one of the parameters
obtained by the EBSD analysis, was used to evaluate the degree of 1 n
m= ∑ P ð1Þ
misorientation in grains. The definition of GAM is as follows. 2 i=1 i
The schematic illustration showing a grain composed of
pixels in the EBSD measurement is shown in Fig. 3. The thick where n is the number of pixels within the grain and Pi the
lines in the figure represent the grain boundaries determined number of pixels neighboring to the pixel i. Therefore, average
as the boundaries which have misorientation larger than 5°. A misorientation β in the grain is defined as,
grain is defined as a region completely surrounded by the
thick boundaries (>5°). It should be noted, here, that the grain 1 m
β= ∑ θ ð2Þ
still can involve misorientation larger than 5° as an isolated m j=1 j
916 MA TE RI A L S CH A R A CT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3–9 2 2

influence of these parameters because they are constant in all


measurements.
The results of the inverse pole figure maps and image
quality maps indicate that it is possible to evaluate the
microstructural change along with the increase of plastic
strain by such maps, but the evaluation can be done only in a
qualitative manner from these maps. Then, in order to
evaluate the change quantitatively, the average misorienta-
tion β determined in Eq. (2) is used. The average misorienta-
tion maps for the specimens strained to 0%, 0.87% (εpl = 0.0087),
2.82% (εpl = 0.0278), 4.79% (εpl = 0.0468) and 9.70% (εpl = 0.0926)
are shown in Fig. 4(c). The β of the undeformed (0%) specimen
is not zero, which is probably due to an error of measurement.
It is thought that such deviations are inevitable in ordinary
measurement conditions. These maps show that β increases
with increasing strain, and the change in misorientation is
different depending on the grains. The same tendency was
Fig. 3 – Schematic illustration showing a grain composed of obtained in the results for the compression specimens.
pixels in the EBSD measurement. The relationship between GAM and plastic strain is shown
in Fig. 5. A unique linear correlation can be obtained between
the GAM and both tensile and compressive strain as:

GAM = 4:373εpl + 0:252: ð4Þ


Here, θj is each misorientation value. The parameter GAM that
shows the average misorientation for the whole analyzed area The maximum and minimum values of β (the average
is defined by the following equation. misorientation of each grain determined in Eq. (2)) in each
  plastic strain are also shown in the figure. It should be noted
=∑
N N
GAM = ∑ ðβk Ak Þ Ak ð3Þ that the bars in Fig. 5 do not indicate standard deviations. Both
k=1 k=1 of the maximum and minimum β increase with the increase of
where N is the number of grains included in the analyzed area plastic strain. This means that the GAM can be an appropriate
and Ak the area of grain k. parameter to evaluate the amount of plastic strain. At the
same time, the difference between the maximum and the
minimum values expands with increasing strain, which
3. Results indicates that the change in misorientation within the grains
is not uniform but heterogeneous in grain to grain. The
3.1. Relation between the EBSD Parameter and Plastic coefficients in Eq. (4) would vary depending on the measure-
Strain by Tension and Compression ment conditions. However it seems that this linear correlation
is a universal tendency as far as the conditions are controlled
The EBSD results for the tensile specimens strained to 0%, well.
0.87% (εpl = 0.0087), 2.82% (εpl = 0.0278), 4.79% (εpl = 0.0468) and It is summarized that the GAM has a unique linear
9.70% (εpl = 0.0926) are shown in Fig. 4. The inverse pole figure correlation with true plastic strain and can be an appropriate
maps shown in Fig. 4(a) indicate the crystal orientations parameter to evaluate plastic deformations quantitatively. It
parallel to the normal direction (ND) of the observed planes. should be noted, however, that the analyzed area has to
These maps show that the degree of color gradation within the include a sufficient number of grains, when the changes in
grains increases with plastic strain. It means that the misorientation are heterogeneous in grain to grain. Further-
orientations of the grains are initially uniform but misorien- more, it has been shown in this study that the relationship
tation appears within the grains as the plastic deformation between GAM and true plastic strain is independent of the
proceeds. The image quality maps shown in Fig. 4(b) indicate direction of deformation, that is, tensile or compression.
the quality of the EBSD patterns in gray scale. It is known that
higher strained region has lower value of image quality [14]. 3.2. Relation between the EBSD Parameter and Creep
The image quality lowers with increasing strain. However, the Damage
change in image quality within each grain is not uniform but
heterogeneous. Such heterogeneity within the grains is very The results of creep tests have been already shown in Fig. 2.
similar to the color gradation observed in the inverse pole The total engineering strain (e) in creep curve shown in Fig. 2
figure maps. It means that the distribution of lattice defects (b) includes instantaneous elastic strain εe, instantaneous
such as dislocations has a relationship with the change in plastic strain εpl, and creep strain εc. The εc is the time-
crystal orientation in the grains. As is well-known, the image dependent strain component. In this study, the relationship
quality can be influenced not only by strain within the between inelastic strain (εin) that is expressed as, εin = εpl + ε̇c ⋅t,
material, but also by SEM or EBSD acquisition parameters. and EBSD parameters are discussed. The creep strain (εc)
However, in this study, it can be considered that there is no occupies 37–72% of the inelastic strain (εin) for creep damage
M A TE RI A L S CH A RACT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3 –9 2 2 917

Fig. 4 – EBSD results for the type 316NG stainless steel tensile tested by 0–9.26% strain : (a) inverse pole figure maps showing the
crystal orientations parallel to ND of the observed planes, (b) image quality maps and (c) average misorientation maps.

ratios, Dc = 0.12–0.7. The plastic strain εpl of 4.5% estimated and creep deformations are different. However, we discuss it
by using the stress–strain curve obtained in parallel exami- without partitioning the inelastic strain into two components
nation is a considerable amount in inelastic strain (εin), though hereafter.
elastic strain εe of 0.17% calculated from Young's modulus The microstructures of the interruption and rupture speci-
and stress is negligibly small under the applied stress of mens observed by SEM are shown in Fig. 6. Creep voids are
270 MPa. Strictly speaking, EBSD parameters should be related found in the specimens above Dc = 0.50. However there are not
with the plastic strain and the time-dependent creep strain so many voids even in the specimen at Dc = 0.70, which
individually, because deformation mechanisms for uniaxial corresponds to the beginning of tertiary creep. Therefore, it
can be concluded that there is no elongation caused by void
growth below Dc = 0.70.
The EBSD analysis was carried out on the specimens tested
under the applied stress of 270 MPa at 600 °C for various Dc.
The specimen of Dc = 0 held at 600 °C for 599 h without applied
stress was also evaluated by EBSD for comparison. The inverse
pole figure maps showing the crystal orientations parallel to
ND of the observed planes for these specimens are shown in
Fig. 7. There is no color gradation in the crystal orientation
map for the specimen of Dc = 0, but color gradation appears
within the grains after Dc = 0.12. The degree of the gradation
becomes large with the increase in creep damage. It means
that the orientations of the grains are initially uniform but
misorientation appears within the grains as the creep damage
proceeds, as in the case of tensile/compression deformations.
The image quality maps for the creep specimens are shown
in Fig.8. The quality of EBSD patterns lowers with increasing
the creep damage ratio, and especially the ruptured specimen
Fig. 5 – Change in the grain average misorientation (GAM) reveals obviously low image quality. However, the image
with plastic strain in the tensile and compression specimens quality of each grain varies widely in the region of small
of the type 316NG. damage ratio because the effect of crystal orientation is
918 MA TE RI A L S CH A R A CT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3–9 2 2

Fig. 6 – SEM microstructures showing creep voids in the type 316 stainless steel creep tested at 600 °C for various periods.
(a) Creep damage ratio, Dc = 0.12 (70 h), (b) Dc = 0.24 (140 h), (c) Dc = 0.35 (210 h), (d) Dc = 0.50 (300 h), (e) Dc = 0.70 (420 h) and
(f) rupture (599 h).

Fig. 7 – Inverse pole figure maps showing the crystal orientations parallel to ND of the observed planes obtained from the EBSD
measurements for the type 316 stainless steel creep tested at 600 °C for various periods : (a) Dc = 0, (b) Dc = 0.12 (70 h), (c) Dc = 0.24
(140 h), (d) Dc = 0.35 (210 h), (e) Dc = 0.50 (300 h), (f) Dc = 0.70 (420 h) and (g) rupture (599 h).
M A TE RI A L S CH A RACT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3 –9 2 2 919

Fig. 8 – Image quality maps obtained from the EBSD measurements for the type 316 stainless steel creep tested at 600 °C for
various periods : (a) Dc = 0, (b) Dc = 0.12 (70 h), (c) Dc = 0.24 (140 h), (d) Dc = 0.35 (210 h), (e) Dc = 0.50 (300 h), (f) Dc = 0.70 (420 h) and (g)
rupture (599 h).

relatively large as in the case of tensile and compression specimens. The tendency of the change in β and GAM is very
deformation, which was discussed in Section 3.1. The similar to than in the case of plastic deformation. However, a
quantitative change of image quality is described later. sharp rise at the beginning of creep caused by instantaneous
Therefore, it seems that the image quality cannot be a plastic strain is characteristic.
parameter indicating creep damage.
The average misorientation maps for the creep specimens
are shown in Fig. 9. It is shown that the misorientation 4. Discussions about the Parameters for
increases with increasing the creep damage ratio. This Damage Evaluation
tendency is very similar to those in tensile and compression
deformation. The misorientation largely differs depending on In Fig. 11, the change in GAM with the increase of creep
the grains, which indicates that the non-uniform creep damage is compared with the changes in the image quality,
deformation is progressing in the specimen. In addition, which is another quantitative parameter obtained by the EBSD
some voids exist at grain boundaries having large misorienta- analysis, and Vickers hardness, which has been often used as
tions in the specimen of creep damage ratio Dc = 0.7 and the a conventional damage parameter. The image quality of the
ruptured specimen. Accordingly, it is concluded that the ruptured specimen of Dc = 1 is obviously lower than the initial
highly creep damaged area can be determined by the average specimen of Dc = 0. However, the image quality of the
misorientation maps and the average misorientation seems to interrupted specimens of Dc = 0.12–0.70 does not decrease
be an appropriate parameter indicating creep damage. with creep damage ratio. Therefore, the image quality is not
The relationship between GAM and creep damage ratio is an appropriate parameter to evaluate the creep damage. The
shown in Fig. 10. GAM means the average misorientation for hardness value dramatically changes at the instantaneous
the whole analyzed area, as described in Eq. (3). The change in loading or the primary creep, but reveals only a slight increase
GAM is plotted as a function of damage ratio in Fig. 10, at the secondary creep region. Therefore, it is concluded that
together with the minimum and maximum values of the the reliability of the creep life estimation using hardness is
average misorientation (β) within each grain as bars. It is fairly low, especially in the secondary creep region, neverthe-
obvious that both the β of each grain and the GAM of the whole less it is the most important region for the real structure
analysis area monotonously increase as the creep damage components. On the other hand, GAM monotonously
ratio increases. The magnitude of the increase in β with creep increases through all creep stages. After the comparison, it
damage differs depending on the grains. The grains having can be concluded that the GAM is the most suitable parameter
large and small β coexist in the high damaged or ruptured to evaluate creep damage.
920 MA TE RI A L S CH A R A CT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3–9 2 2

Fig. 9 – Average misorientation maps obtained from the EBSD measurements for the type 316 stainless steel creep tested at
600 °C for various periods : (a) Dc = 0, (b) Dc = 0.12 (70 h), (c) Dc = 0.24 (140 h), (d) Dc = 0.35 (210 h), (e) Dc = 0.50 (300 h), (f) Dc = 0.70
(420 h) and (g) rupture (599 h).

Then, the relationship between GAM and strain is dis- macroscopically measured strain includes the strain induced
cussed. The comparison of the change in ΔGAM with by the growth of voids. It means that the ΔGAM measured by
engineering inelastic strain (εin) is shown in Fig. 12. Here, EBSD represents intrinsic deformation, while macroscopic
ΔGAM means the difference of grain average misorientation strain (εin) involves both intrinsic deformation and apparent
from the value of the unstrained specimen. ΔGAM is used one due to voids. Therefore, it can be concluded that ΔGAM is
instead of GAM because GAM is not zero at the unstrained the most meaningful physical parameter even at tertiary creep
specimen. Fig. 12 shows that both curves fairly coincide till the stage.
creep damage ratio of Dc = 0.7 which corresponds to the The relation between ΔGAM and true inelastic strain is
beginning of tertiary creep. The gap between inelastic strain shown in Fig. 13. ΔGAM shows an excellent correlation with
and ΔGAM curves increases after Dc = 0.70. This is because the

Fig. 10 – Change in the grain average misorientation (GAM) Fig. 11 – Change in three kinds of parameters for damage
with creep damage ratio in the specimens of the type 316 evaluation as functions of creep damage ratio in the specimens
stainless steel creep tested at 600 °C. of the type 316 stainless steel creep tested at 600 °C.
M A TE RI A L S CH A RACT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3 –9 2 2 921

5. Conclusions

The EBSD method was applied to the deformation and damage


evaluation of austenitic stainless steels strained by tension or
compression and also tested in creep. It was found that the
parameter, Grain Average Misorientation (GAM) which shows
the average misorientation for the whole observed area, is
useful for quantifying the microstuructural change as the
plastic or creep strain increases. The following results were
obtained in this study.

(1) The linear correlation was obtained between GAM and


plastic strain in tension and compression. By applying
this linear correlation to the materials of which plastic
strain is unknown, the induced plastic strain can be
Fig. 12 – Changes in ΔGAM and engineering inelastic strain as
estimated.
a function of creep damage ratio in the specimens of the type
(2) The GAM increases monotonously as the creep damage
316 stainless steel creep tested at 600 °C.
ratio increases. However, at the tertiary creep stage, the
increasing rate of GAM is smaller than that of the
inelastic strain macroscopically measured in the spec-
imen gage length. This is because the inelastic strain
the inelastic strain below strain of about 0.2 at which the
includes the strain induced by creep void growth.
tertiary creep begins. It seems possible to use this relationship
Therefore, ΔGAM represents the essential deformation
as a master curve for creep life prediction.
behavior within the grain.
The result obtained by tensile and compressive test shown
(3) The present result indicates that the parameter GAM
in Fig. 5 is also plotted in Fig. 13. These two curves in Fig. 13
obtained by EBSD analysis can be applied as one of the
monotonously increase with true inelastic strain (εin), but
useful and meaningful physical parameters for creep
have different gradients. As mentioned in Section 2.4, the step
life prediction.
sizes used in the EBSD analysis for tensile/compression
specimens and creep specimens are 2 and 0.8 μm, respective-
ly. The difference in step size is thought to be one of the main
REFERENCES
causes for different gradient in ΔGAM–inelastic strain relation.
The effect of step size on misorientation is a very important
issue which must be discussed in the future study. When the [1] Edited by The Japan Society of Mechanical Engineers.
effect of step size on the ΔGAM–inelastic strain relation is Remaining Life Estimation Technology for Power Plants and
established, the creep damage can be estimated only from the Structures. Tokyo: Gihodo Shuppan Co. Ltd.; 1992 (in
Japanese).
results of simple tests in a short time, such as a tensile or
[2] Wilkinson AJ. Measurement of elastic strains and small lattice
compression test.
rotations using electron back scatter diffraction.
Ultramicroscopy 1996;62:237–47.
[3] Wilkinson AJ. A new method for determining small
misorientations from electron back scatter diffraction
patterns. Scr Mater 2001;44:2379–85.
[4] MacLachlan DW, Wright LW, Gunturi S, Knowles DM.
Constitutive modelling of anisotropic creep deformation in
single crystal blade alloys SRR99 and CMSX-4. Int J Plast
2001;17:441–67.
[5] Field DP, True BW, Lillo TM, Flinn JE. Observation of twin
boundary migration in copper during deformation. Mater Sci
Eng A 2004;372:173–9.
[6] Boyce BI, Michael JR, Kotula PG. Fatigue of metallic
microdevices and the role of fatigue-induced surface oxides.
Acta Mater 2004;52:1609–19.
[7] Field DP, Trivedi PB, Wright SI, Kumar M. Analysis of local
orientation gradients in deformed single crystals.
Ultramicroscopy 2005;103:33–9.
[8] Wilkinson AJ, Meaden G, Dingley DJ. High-resolution elastic
strain measurement from electron backscatter diffraction
patterns: new levels of sensitivity. Ultramicroscopy 2006;106:
Fig. 13 – Change in ΔGAM as a function of true inelastic strain
307–13.
in the specimens of the type 316 stainless steel creep tested [9] Trivedi PB, Yassar RS, Field DP, Alldredge R. Microstructural
at 600 °C and the type 316NG stainless steel tensile or evolution and observed stress response during hot deformation
compression tested. of 5005 and 6022 Al alloys. Mater Sci Eng A 2006;425:205–12.
922 MA TE RI A L S CH A R A CT ER IZ A TI O N 61 ( 20 1 0 ) 9 1 3–9 2 2

[10] Pantleon W. Resolving the geometrically necessary dislocation [22] Fukuoka C, Morishima K, Yoshizawa H, Mino K.
content by conventional electron backscattering diffraction. Scr Misorientation development in grains of tensile strained and
Mater 2008;58:994–7. crept 2.25%Cr–1%Mo steel. Scr Mater 2002;46:61–6.
[11] Merriman CC, Field DP, Trivedi P. Orientation dependence of [23] Trivedi P, Field DP, Weiland H. Alloying effects on dislocation
dislocation structure evolution during cold rolling of substructure evolution of aluminum alloys. Int J Plast 2004;20:
aluminum. Mater Sci Eng A 2008;494:28–35. 459–76.
[12] He W, Ma W, Pantleon W. Microstructure of individual grains [24] Pantleon W. On the apparent saturation of the average
in cold-rolled aluminium from orientation inhomogeneities disorientation angle with plastic deformation. Scr Mater
resolved by electron backscattering diffraction. Mater Sci Eng 2005;53:757–62.
A 2008;494:21–7. [25] Kamaya M, Wilkinson AJ, Titchmarsh JM. Measurement of
[13] Wilkinson AJ, Meaden G, Dingley DJ. Mapping strains at the plastic strain of polycrystalline material by electron
nanoscale using electron back scatter diffraction. backscatter diffraction. Nucl Eng Des 2005;235:713–25.
Superlattices Microstruct 2009;45:285–94. [26] Kamaya M, Wilkinson AJ, Titchmarsh JM. Quantification of
[14] Wilkinson AJ, Adams BL. Measuring strains using electron plastic strain of stainless steel and nickel alloy by electron
backscatter diffraction. In: Schwartz AJ, Kumar M, editors. backscatter diffraction. Acta Mater 2006;54:539–48.
Electron Backscatter Diffraction in Material Science. New [27] Kimura H, Wang Y, Akiniwa Y, Tanaka K. Misorientation
York: Kluwer Academic/Plenum Publishers; 2000. p. 231. analysis of plastic deformation of austenitic stainless steel by
[15] Wilkinson AJ, Dingley DJ. Quantitative deformation studies EBSD and X-ray diffraction methods. J Jpn Soc Mech Eng Ser A
using electron back scatter patterns. Acta Metall Mater 2005;71:118–24 (in Japanese).
1991;39:3047–55. [28] Wert JA, Huang X, Winther G, Pantleon W, Poulsen HF.
[16] Wilkinson AJ. Deformation studies of metal matrix Revealing deformation microstructures. Mater Today 2007;10:
composites using electron backscatter patterns. Mater Sci Eng 24–32.
A 1991;135:189–93. [29] Takaku R, Saito D, Yoshioka Y. Effect of grain size and crystal
[17] Wilkinson AJ, Dingley DJ. The distribution of plastic orientation on creep damage evaluation by changes of
deformation in a metal matrix composite caused by straining misorientation in Hastelloy X. J Soc Mater Sci Jpn 2009;58:
transverse to the fibre direction. Acta Metall Mater 1992;40: 229–34 (in Japanese).
3357–68. [30] Fujiyama K, Mori K, Kaneko D, Matsunaga T, Kimachi H. Creep
[18] Wilkinson AJ, Hirsch PB. Electron diffraction based damage assessment of high chromium steel forging through
techniques in scanning electron microscopy of bulk EBSD method and hardness measurement. J Jpn Soc Mech Eng
materials. Micron 1997;28:279–308. Ser A 2008;74:323–8 in Japanese.
[19] Sutliff JA. An investigation of plastic strain in copper by [31] Ohtani T, Yin F, Kamada Y. Microstructural and
automated-EBSP. Microsc Microanal 1999;5(Suppl 2):236. non-destructive evaluation of creep damage in martensitic
[20] Lehockey EM, Lin YP, Lepik OE. Mapping residual plastic strain stainless steel. J Soc Mater Sci Jpn 2009;58:136–42 (in
in materials using electron backscatter diffraction. In: Japanese).
Schwartz AJ, Kumar M, Adams BL, editors. Electron [32] Mitsuhara M, Morioka S, Hata S, Ikeda K, Nakashima H.
Backscatter Diffraction in Material Science. New York: Kluwer Degradation behavior of crept lath martensite analyzed by
Academic/Plenum Publishers; 2000. p. 247. crystallographic evolution. Report of the 123rd committee on
[21] Othon MA, Brewer LN, Angeliu TM, Young LM. Electron heat-resisting metals and alloys. Jpn Soc Promot Sci 2009;50:
back-scattered diffraction misorientation mapping applied to 37–43 in Japanese.
stress corrosion cracking of stainless steels. Microsc
Microanal 2002;8(Suppl 2):698CD.

You might also like