Está en la página 1de 7

Effect of microstructure on hydrogen embrittlement of various

stainless steels
C.L. Lai
a
, L.W. Tsay
b,n
, C. Chen
c
a
Institute of Nuclear Energy Research (INER), Taoyuan 325, Taiwan
b
Institute of Materials Engineering, National Taiwan Ocean University, Keelung 202, Taiwan
c
Department of Material Science and Engineering, National Taiwan University, Taipei 106, Taiwan
a r t i c l e i n f o
Article history:
Received 1 November 2012
Received in revised form
1 July 2013
Accepted 4 July 2013
Available online 10 July 2013
Keywords:
Stainless steels
Hydrogen embrittlement
Microstructure
Strain-induced
-martensite
a b s t r a c t
This work investigated the effect of microstructure on the susceptibility of 304L (metastable), 310S
(stable) austenitic and 410 martensitic stainless steels (SSs) to hydrogen embrittlement (HE). Slow-
displacement-rate notched tensile tests were performed at various combinations of temperature (25 and
80 1C) and environment (air and H
2
) to evaluate the relative HE susceptibility of these alloys. At 25 1C, the
untempered 410 SS was the specimen most susceptible to HE among the investigated specimens,
whereas the 310S and tempered 410 specimens exhibited low HE susceptibility. The formation of strain-
induced -martensite in a localized region in front of the notch tip was the main cause for the high HE
susceptibility of the 304L SS tested at 25 1C. In general, the HE susceptibility was reduced to various
degrees for specimens tested at 80 1C. A signicantly lower susceptibility to HE was observed for the
304L specimen at 80 1C due to the suppressed formation of -martensite in the highly strained region.
& 2013 Elsevier B.V. All rights reserved.
1. Introduction
Austenitic stainless steels (SSs), which possess high corrosion
resistance and good mechanical properties, are widely used in piping
systems in the petrochemical industry and in nuclear power plants
[1,2]. They are also considered candidates for hydrogen storage
materials [3]. Depending on their austenite stability [48], metastable
austenitic SSs may undergo phase transformations from austenite to
ferromagnetic and paramagnetic martensites during plastic
deformation. Angel [9] suggested the use of the Md
30
temperature,
i.e., the temperature at which 50% of the austenite phase transforms
into martensite during tensile testing at a true strain of 0.3, to
evaluate the stability of austenitic SSs. Higher Md
30
temperatures
clearly result in less stable alloys. The formula that relates the Md
30
to
the chemical composition of an alloy is as follows:
Md
30
1C 413462%C %N9:2%Si
8:1%Mn13:7%Cr9:5%Ni18:5%Mo
The inuence of strain-induced on hydrogen embrittlement
(HE) in several austenitic SSs has been reported extensively [1013].
The diffusion coefcient of hydrogen in -martensite (bct) is higher
than that in (fcc) and -martensite (hcp) [14]. Strain-induced
-martensite is reported to act as a hydrogen diffusion highway in
hydrogen-charged 304 SS [15], resulting in an increased hydrogen
concentration at the crack tip and accelerated fatigue crack growth.
Furthermore, sensitization causes the transformation of -martensite
to occur preferentially along the grain boundaries of 304 and 316 SSs,
providing a short-circuit path for hydrogen transportation to the crack
tip [16] and enhances the intergranular fracture and HE susceptibility
of the material [17]. In the case of a stable 310S SS, no strain-induced
-martensite occurs, and HE can only take place when the hydrogen
concentration in front of the crack tip exceeds a certain critical value
[18]. In addition to the formation of -martensite, the rate of
hydrogen transport, the stacking fault energy (SFE) and the test
temperature can affect the HE susceptibility of alloys [1921]. In this
study, the effects of microstructure and temperature on the relative
HE susceptibility of 304L, 310S and 410 SS (untempered and tem-
pered) specimens in gaseous hydrogen were investigated. The HE
susceptibility of these SSs was correlated with the corresponding
microstructures, particularly those of the metastable 304L specimens,
at different temperatures.
2. Experimental procedures
The materials used in this study included AISI 304L, 310S and
410 SSs. The 304L and 310S specimens were tested in the solution-
annealed condition, whereas the 410 specimens were tested in
untempered or air-hardened (410A, 980 1C/1 h+air-cooled) and
tempered (410T, 980 1C/1 h+air-cooled+700 1C/2 h) conditions.
The chemical compositions of the materials used are listed in
Table 1. Notched tensile tests were conducted on a screw-driven
Contents lists available at SciVerse ScienceDirect
journal homepage: www.elsevier.com/locate/msea
Materials Science & Engineering A
0921-5093/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2013.07.004
n
Corresponding author. Tel.: +886 224622192; fax: +886 224635730.
E-mail address: b0186@mail.ntou.edu.tw (L.W. Tsay).
Materials Science & Engineering A 584 (2013) 1420
machine (GOTECH, model: GT-7007-MC10) at a crosshead speed of
1.210
4
mm/s. The precise speed control of the displacement was
achieved by an AC servo motor equipped with an accurate gear box.
The applied load (determined by a load cell) and displacement
(measured by a linear variable differential transformer) of the
machine were calibrated prior to testing. The test conditions included
different combinations of temperature (25 and 80 1C) and environ-
ment (air and H
2
). For the specimens tested in hydrogen, a stainless
steel chamber with a controlled hydrogen pressure was employed.
After installing the specimen, high-purity (99.99%) hydrogen gas was
introduced into the chamber at a pressure of 2 MPa. Fig. 1 shows a
double-edge notched tensile specimen with a thickness of 3 mm,
in which two sharp notches conne the region in which phase
transformation, if any occurs, takes place. The content of strain-
induced -martensite in front of the notch tips of the 304L specimens
was measured using a Ferrite scope (Fischer Feritscope MP30) during
the tests in air at 25 and 80 1C. For the specimens tested in gaseous
hydrogen, the content was measured only after the tests. The
amount of ferrite, as measured using the Ferrite scope, represented
the average -martensite content yielding from a certain volume at
the detected site of the specimen [22]. Although the -martensite
content might change with the distance from the detected site, the
relative amounts of ferrite among the specimens could still be
compared. For convenience, the specimens were designated accord-
ing to the conditions under which they were tested, e.g., the 310S
specimen tested in air at 25 1C and the 304L specimen tested
in hydrogen at 80 1C were named the 310S-25a and 304L-80h speci-
mens, respectively. For the specimens tested in H
2
, the loss of notch
tensile strength (NTS) was used as an index of the HE susceptibility,
i.e., higher NTS losses indicated higher HE susceptibilities for the
specimens. The formula used to calculate the NTS loss is as follows:
NTS loss % NTS in airNTS in H
2
=NTS in air 100%
Fracture appearance in the specimens after the notched tensile
tests was examined by using an LEO-1530 eld emission scanning
electron microscope (FE-SEM). The specimens used in the electron
backscattered diffraction (EBSD) analyses were electropolished
with a solution containing 5% perchloric acid+95% acetic acid.
The EBSD system (EDAX-TSL orientation imaging microscopy)
attached to the FE-SEM was used to distinguish the phase
constituents in the strained regions near cracks. In addition, the
relationship between the fracture path and the underlying micro-
structure was also investigated.
3. Results and discussion
3.1. Microstructure and microhardness
Fig. 2 shows photographs that illustrate the microstructures of
various specimens. The solution-annealed 304L and 310S speci-
mens consisted of equiaxed austenite with annealing twins as
shown in Fig. 2(a) and (b), respectively. The hardnesses of the 304L
and 310S specimens were similar: approximately HV 167 for the
former and HV 157 for the latter. In contrast, the 410A specimen
was composed primarily of untempered martensite (HV 400), as
shown in Fig. 2(c). To obtain proper combinations of toughness
and strength, air-hardened 410 SS is often tempered at 700 1C for
2 h. Fig. 2(d) shows a typical tempered martensite structure in the
410T specimen with a lower hardness of HV 265. The detailed
microstructures of the 410A and 410T specimens are also shown in
the right halves of Fig. 2(c) and (d), respectively. The formation of
lath martensite, which has a high dislocation density, accounted
for the high hardness of the 410A specimen (Fig. 2(c)). After
tempering the specimen at 700 1C for 2 h, the loss of lath
morphology associated with reduced dislocation density and the
precipitation of M
23
C
6
carbides resulted in a lower hardness value
for the 410T specimen (Fig. 2(d)).
3.2. Notched tensile test
The Md
30
temperatures of the 304L and 310S specimens were
calculated, based on the compositions of the specimens, to be 62 and
146 1C, respectively. Unlike the metastable 304L specimen, the 310S
specimen was stable, and strain-induced -martensite likely did not
form in front of the notch tips during tensile straining. Fig. 3 displays
the variation in NTS and the change in the content of the 304L
specimens tested at 25 and 80 1C during straining. The ferritescope
measurements of the content were performed only at the notch
tips of the specimens during testing. At the test temperature of 25 1C,
the formation of -martensite was induced extensively at the notch
front after an applied stress of approximately 400 MPa was reached.
Beyond this point, a rapid increase in the content, from 0% to 43.7%
at the end of the test, was observed. When the test temperature was
increased to 80 1C, the content was measurable only after the
maximum strength or NTS was attained. The amount of (7.8%)
measured at the end of the notched tensile test at 80 1C was
considerably lower than that measured at 25 1C (43.7%). This result
implies that the formation of was substantially suppressed by the
increase in the testing temperature. Compared with the 304L speci-
mens tested in air, the 304L-25h and 304L-80h specimens exhibited
signicant decreases, 6.7% and 1.9%, respectively, in their contents
at fracture.
Fig. 4 shows front-view photographs of the fractured 304L-25a
and 304L-25h specimens. Visual observations revealed that the plastic
zones ahead of the notch tips in the 304L-25a specimen were
approximately 2 mm in radius (Fig. 4(a)). In contrast, the size of the
plastic zones in the 304L-25h specimen was difcult to visualize due
to the restricted deformation in hydrogen (Fig. 4(b)). Parallel cracks
could be observed on the side surfaces of 304L-25h specimen. The
premature failure of the 304L-25h specimen was related to HE
associated with crack formation in small deformation zones, resulting
in the lack of gross plastic deformation ahead of the notch tips. Fig. 5
Table 1
The chemical compositions (wt%) of the alloys used in this study.
Meterials Elements
C Cr Ni Mn Si P S Mo Cu Fe
304L 0.019 18.20 8.04 1.53 0.50 0.030 0.003 0.40 Bal.
310S 0.05 24.66 19.09 1.17 0.80 0.015 0.003 Bal.
410 0.13 12.18 0.13 0.16 0.35 0.013 0.003 0.023 0.016 Bal.
Unit : mm
Fig. 1. Schematic diagram showing the dimensions of the notched tensile
specimens.
C.L. Lai et al. / Materials Science & Engineering A 584 (2013) 1420 15
shows the appearance of several specimens after notched tensile
tests, with the focus on only one of the notch roots due to symmetry.
For the specimens that were insensitive to HE, e.g., the 310S SS,
similar plastic zones near the notch tips were observed in the
specimens tested in air and in hydrogen (Fig. 5(a)). In the case of
the 410A-25a specimen tested in air, the inherent high strength/low
ductility resulted in a very small or negligible plastic zone ahead of
the notch front even in air (shown in the left half of Fig. 5(b)). As
expected, this specimen was susceptible to HE, due to the nature of
untempered martensite, and exhibited the brittle fracture with no
plasticity (the right half of Fig. 5(b)). After tempering at 700 1C for 2 h,
the lowered strength and improved ductility of the 410T-25a speci-
men showed an increase in plasticity and a moderate increase in its
plastic zone (the left half of Fig. 5(c)) relative to the 410A-25a
specimen. Furthermore, the plastic zone size of the 410T-25h
50 m 50 m
0.2 m
OM
20 m
TEM
0.2 m 20 m
OM TEM
Fig. 2. Microstructures of various specimens: (a) 304L; (b) 310S; (c) 410A and (d) 410T.
Fig. 3. Change in the content and the notched tensile strength plotted against
the displacement curves of the 304L specimens tested in air. Note that the
crosshead speed used in the tests was 6 10
4
mm/s and all the content
measurements were performed at the notch tips during testing.
2 mm
2 mm
Fig. 4. Front views of the fractured specimens: (a) 304L-25a and (b) 304L-25h.
C.L. Lai et al. / Materials Science & Engineering A 584 (2013) 1420 16
specimen (the right half of Fig. 5(c)) was smaller than that of the
410T-25a specimen as a result of HE.
Notched tensile strength versus displacement curves for all of
the specimens tested in air (solid lines) and H
2
(dotted lines) at
different temperatures are plotted in Fig. 6. The results indicate
that the 310S specimen exhibited a minor change in strength and
displacement with increasing temperature and showed low HE
susceptibility when tested in gaseous hydrogen at both 25 and
80 1C. The 304L-25a specimen exhibited greater strength and
deformation than the 304L-80a specimen. These differences can
be attributed to the induced formation of -martensite in the
304L-25a specimen, which strengthened the material and resulted
in a delay in localized necking during testing.
As Fig. 6(a) shows, the 304L-25a and 304L-25h curves overlap in
the displacement of 00.5 mm, suggesting that the deterioration
effect of HE did not occur due to the lack of -martensite ahead of
the notch tips. However, HE associated with crack formation might
have occurred in the 304L-25h specimen during the test in hydrogen
if sufcient -martensite was formed in small deformation zones
ahead of the notch tips. The strength corresponding to the 0.5-mm
displacement, i.e., approximately 230 MPa, could represent a thresh-
old value for HE of the 304L specimens tested at room temperature.
Under such circumstances, the strength of the 304L-25h specimen
was signicantly lower than that of the 304L-25a specimen for
displacements greater than 0.5 mm (Fig. 6(a)).
With respect to the 304L-25h and 304L-80h specimens, HE led
to a signicant reduction in strength and deformability relative to
the corresponding specimens tested in air. In the presence of
hydrogen, the decrease in strength and deformability of the 304L
specimen at 80 1C was considerably lower than that of the speci-
men tested at 25 1C. This phenomenon indicates that the HE of the
304L SS was strongly temperature-dependent, which could be
related to the amount of strain-induced formed during the
notched tensile test. The high resistance to HE of the 310S SS,
compared to that of the 304L SS, could be attributed to the high
stability of austenite and is discussed later in the text. Among the
specimens tested in air, the 410A specimen displayed high
Fig. 5. Front views of the fractured specimens tested at 25 1C: (a) 310S; (b) 410A;
and (c) 410T.
a
h
a
h
a
h
A-80a
A-80h
a
h
a
h
a
h
A-25a
A-25h
Fig. 6. Notched tensile strength vs. displacement curves of various SS specimens
tested at (a) 25 1C and (b) 80 1C.
C.L. Lai et al. / Materials Science & Engineering A 584 (2013) 1420 17
strength and low ductility, irrespective of the test temperature.
However, a sharp decrease in strength and ductility, which is
indicative of a material suffering from severe HE, was observed for
the 410A specimen tested in H
2
(Fig. 6). The results also show that
the 410T specimen (tempered at 700 1C for 2 h) exhibited a
reduced susceptibility to HE.
Fig. 7 shows the NTS loss of various SS specimens tested at 25 and
80 1C. The 410A specimen displayed the greatest NTS loss (67.5% at
25 1C and 57.6% at 80 1C). These results reveal that the increase in
temperature had only a slight effect on reducing the HE susceptibility
of the 410A specimen. However, the 310S and 410T specimens
showed low susceptibility to gaseous HE (NTS loss less than 6%) at
both test temperatures. Notably, the 304L specimen was highly
susceptible to HE, with an NTS loss of 34.5% at 25 1C, but the NTS
loss was reduced to 8.4% at 80 1C. At 80 1C, the NTS loss of most of
the specimens was low, except for that of the 410A specimen. It is
obvious that the NTS loss or HE susceptibility of the 304L specimen
(a metastable SS) was strongly affected by the temperature.
3.3. Hydrogen embrittlement mechanism
Birnbaum et al. [23] have explained the loss of gross ductility in
the presence of hydrogen based on hydrogen-enhanced localized
plasticity (HELP). This theory is based on hydrogen atoms not only
enhancing dislocation motion but also resulting in localized plastic
deformation, which is, in turn, responsible for reduced macroscopic
ductility [2426]. Hydrogen is known to move toward the plastic zone
ahead of a crack tip, through mobile dislocations or by diffusion, to
embrittle steels. The hydrogen diffusivity of solution-annealed auste-
nitic SSs is reported to be approximately 2.803.9610
14
m
2
/s at
110 1C [15] and 1.15.510
15
m
2
/s at 50 1C [27]. In a pre-strained
304 SS that contained approximately 70% -martensite, the hydrogen
diffusivity was found to be about 2.410
13
m
2
/s at 80 1C [27].
Accordingly, the diffusion distance of hydrogen in a highly strained
metastable austenitic SS was estimated to be 5.010
4
mm in 1 s at
80 1C. On the other hand, in studies on the fatigue crack growth rates
of 304 and 316 SSs in hydrogen at room temperature [28,29],
accelerated crack growth could be observed even at a high growth
rate of approximately 10
3
mm/cycle or 210
2
mm/s. It has also
been demonstrated that a thin layer of -martensite is formed on the
fatigue fracture surfaces, accounting for the assisted crack growth of
304L in hydrogen [30]. In the reported study, the depth of the
hydrogen-affected zone during crack propagation was greater than
the depth of penetration by diffusion. It is clear that hydrogen atoms
were transported to highly strained regions predominantly through
mobile dislocations, not by diffusion.
Without the formation of strain-induced martensite, the stable
austenite matrix of the 310S specimen possessed a large plastic zone,
as well as a high solubility for hydrogen. Consequently, a critical
hydrogen concentration in the plastic zone was difcult to attain in
the 310S specimen, making the 310S specimen immune to HE.
Moreover, an improved resistance to HE is associated with an increase
in the fraction and stability of -martensite in modied austenitic SSs
[31,32]. This implies that the extent and amount of -martensite
plays an important role in facilitating HE and dominates the fracture
process of the metastable austenitic SS in hydrogen. Fig. 8(a) shows
the relative size (not to scale) of the plastic zones of metastable 304L
specimens tested in hydrogen and air at low temperatures, e.g., room
temperature. The presence of hydrogen not only enhanced the
localization of the deformation but also led to more strain-induced
-martensite in the 304L-25h specimen, which was an auto-catalytic
process. As a result, the plastic zone size of the 304L-25h specimen
was considerably smaller than that of the 304L-25a specimen, leading
to the 304L-25h specimen being susceptible to HE and exhibiting
high NTS loss at 25 1C. Fig. 8(b) displays a growing crack path in the
304L-25h specimen, in which the plastic zone at the crack tip consists
of high volume fractions of -martensite. A sufciently high fraction
of -martensite is necessary to cause HE and cracking for metastable
austenitic SSs at low temperatures. When the test temperature was
increased to 80 1C, the formation of -martensite was signicantly
suppressed in the 304L-80h specimen, resulting in reduced HE
susceptibility.
To conrm the previously mentioned mechanism of HE, the
notched tensile test was stopped at approximately 95% of the NTS
value of the 304L-25h and 304L-80h specimens. SEM and EBSD
analyses were performed on these specimens, focusing on the region
in front of the notch tip. Fig. 9 reveals that a ne crack that tended to
0 10 20 30 40 50 60 70 80 90 100
0
10
20
30
40
50
60
70
80
90
100
304L
310S
410A
410T
Test temperature (C)
N
T
S

l
o
s
s

(
%
)
Fig. 7. NTS loss of various SS specimens tested at different temperatures.
A previous deformation zone when
the crack tip was located at A
: Strain-induced '
: Cracks
: (matrix)
Crack propagation direction
A
or slip bands
: martensite or slip bands
Plastic zone tested in Air
Plastic zone tested in H2
Crack tip
: Plastic zone
: Dislocation
: Slip line
: H moving path
: H atom
Fig. 8. Schematic diagrams showing (a) the relative sizes of the plastic zones of the
304L-25h and 304L-25a specimens and (b) the trajectories of the plastic zones of a
growing crack in the 304L-25h specimen.
C.L. Lai et al. / Materials Science & Engineering A 584 (2013) 1420 18
propagate along a thin layer of strain-induced -martensite, which
formed layers measuring less than 100 m in thickness on each side
of the fracture plane (Fig. 9(a) and (b)). At 80 1C, the reduced
content and hydrogen adsorption [33] were expected to alleviate the
HE susceptibility of the 304L-80h specimen. Observations of the crack
path at the notch tip revealed the occurrence of plastic deformation in
the 304L-80h specimen prior to the formation of a short crack (Fig. 9
(c)), in contrast to little deformation being observed in the 304L-25h
specimen (Fig. 9(a)). The EBSD map of the 304L-80h specimen (Fig. 9
(d)) reveals that -martensite was formed predominantly around the
crack. The improved ductility of -marten site [31,32] was also helpful
in lowering the HE susceptibility of the 304L-80h specimen. For the
410A specimen with the hardened structure (untempered martensite)
tested in H
2
, a very small or negligible plastic zone (associated with
high strength/hardness) in front of the notch tip facilitated the
hydrogen concentration easily achieving a critical value. Accordingly,
the 410A specimen was susceptible to HE and exhibited high NTS loss,
regardless of the testing temperature. In contrast, the 410T specimen,
consisting of ferrite and carbides, had a lowered strength and an
improved ductility, leading to a reduced susceptibility to HE. Further-
more, a large quantity of carbides (innocuous hydrogen traps) were
expected to retard hydrogen diffusion inward, lowering the HE
susceptibility of the 410T specimen. As shown in Fig. 6, the strength
vs. displacement curves at 80 1C for the 310S and 410 specimens did
not change signicantly from those at 25 1C. These observations are
consistent with the absence of apparent changes in the plastic zone
size for these specimens, tested at either 25 or 80 1C.
3.4. Fractographic examinations
SEM fractographs of various SS specimens after notched tensile
tests are displayed in Fig. 10. All of the specimens tested in air
revealed a ductile fracture mode, which is similar to the fracture
appearance of the 304L-25a specimen (Fig. 10(a)). In contrast, a
quasi-cleavage fracture with secondary cracks was observed in the
304L-25h specimen (Fig. 10(b)), which could be related to the
brittle nature of martensite. The 310S-25h specimen with less
susceptibility to HE also exhibited ductile dimples (Fig. 10(c)). In
the case of the 410A-25h specimen, the high NTS loss was
associated with extensive transgranular brittle fractures with
cracking along prior austenite grain boundaries (Fig. 10(d)). Never-
theless, the 410T-25h specimen with low NTS loss showed trans-
granular fracture inter-dispersed with ductile dimples (Fig. 10(e)).
Overall, the fracture features of the 410T-25h specimen still
showed a somewhat brittle nature. At a lower displacement rate
or with an increased hydrogen supply, the 410T-25h specimen
might exhibit a change in fracture mode. When the test tempera-
ture was increased to 80 1C, the HE susceptibility or the NTS loss of
the 304L-80h specimen decreased, and the extent of the brittle
fracture feature was greatly reduced (Fig. 10(f)). The ductility of the
304L-80h specimen was observed to be lower than that of the
counterpart specimen tested in air (Fig. 6(b)), indicating that the
304L-80h specimen was moderately affected by HE even at 80 1C.
As previously mentioned, the HE susceptibility and the fracture
mode of the specimens that did not undergo phase transformation
during straining (the 310S, 410A and 410T specimens) were
affected by the test temperature to a lesser extent.
4. Conclusions
1. At room temperature, the 410A specimen showed the highest
HE susceptibility, whereas the 310S and tempered 410 speci-
mens exhibited low susceptibility to HE. When the test tem-
perature was increased to 80 1C, the NTS loss of most
specimens was low, except for that of the 410A specimen. The
HE susceptibility of 410 SS could be signicantly reduced by
tempering at the expense of strength.
2. The formation of strain-induced -martensite ahead of the notch
tips was the main reason for the high susceptibility to HE of the
304L SS at 25 1C. Moreover, the suppression of the -to-
transformation during straining at 80 1C led to a large decrease
in the HE susceptibility of 304L SS. In contrast, the 310S SS
specimen did not form the phase during plastic deformation
and was immune to HE.
3. During the straining of 304L SS at 25 1C in H
2
, the premature
cracking of the -martensite in the small deformation zone
ahead of the crack tip led to the high HE susceptibility of the
specimen. Moreover, the transportation of hydrogen by mobile
dislocations to a localized region caused severe embrittle-
ment and accounted for crack propagation connement in the
thin region of -martensite.
100 m
Note that: main crack; ; '; 50 m
50 m
Note that: main crack; ; ';
10 m
Fig. 9. SEM/EBSD analyses near the main crack of the 304L specimens: (a) SEM
image of the 304L-25h specimen; (b) phase identication of the dashed frame in
(a); (c) SEM image of the 304L-80h specimen and (d) phase identication of the
dashed frame in (c). Note that the average condence indices of Fig. 7(b) and (d) are
0.55 and 0.47, respectively.
C.L. Lai et al. / Materials Science & Engineering A 584 (2013) 1420 19
Acknowledgments
The authors gratefully acknowledge the partial support of this
study by the National Science Council of the Republic of China
(NSC 99-2221-E-002-063-MY3 and 100-NU-E-002-002-NU).
References
[1] C. Guerre, O. Raquet, E. Herms, M.L. Calvar, G. Turluer, Proceedings of the 12th
International Conference on Environmental Degradation of Materials in
Nuclear Power System, TMS, 2005, pp. 1029.
[2] M.C. Young, J.Y. Huang, R.C. Kuo, Mater. Trans. 50 (2009) 657.
[3] L. Briottet, I. Moro, P. Lemoine, Int. J. Hydrogen Energy 37 (2012) 17616.
[4] N. Gey, B. Petit, M. Humbert, Metall. Mater. Trans. A 36 (2005) 3291.
[5] B. Petit, N. Gey, M. Cherkaoui, B. Bolle, M. Humbert, Int. J. Plast. 23 (2007) 323.
[6] K. Spencer, J.D. Embury, K.T. Conlon, M. Veron, Y. Brechet, Mater. Sci. Eng.
A387389 (2004) 873.
[7] A. Das, S. Sivaprasad, M. Ghosh, P.C. Chakraborti, S. Tarafder, Mater. Sci. Eng. A
486 (2008) 283.
[8] T. Fukuda, T. Kakeshita, K. Kindo, Mater. Sci. Eng. A438440 (2006) 212.
[9] T. Angel, J. Iron Steel Inst. 177 (1954) 165.
[10] N. Narita, H.K. Birnbaum, Scr. Metall. 14 (1980) 1355.
[11] C.L. Briant, Metall. Trans. A 10 (1979) 181.
[12] D. Eliezer, D.G. Chakrapani, C.J. Altstetter, E.N. Pugh, Metall. Trans. A 10 (1979)
935.
[13] P. Muraleedharan, H.S. Khatak, J.B. Gnanamoorthy, P. Rodriguez, Metall. Trans.
A 16 (1985) 285.
[14] T.P. Perng, C.J. Altstetter, Metall. Trans. A 18 (1987) 123.
[15] Y. Mine, C. Narazaki, K. Murakami, S. Matsuoka, Y. Murakami, Int. J. Hydrogen
Energy 34 (2009) 1097.
[16] G. Han, J. He, S. Fukuyama, K. Yokogawa, Acta Mater. 46 (1998) 4559.
[17] W.J. Li, M.C. Young, C.L. Lai, W. Kai, L.W. Tsay, Corros. Sci. 68 (2013) 25.
[18] J.H. Huang, C.J. Altstetter, Metall. Trans. A 22 (1991) 2605.
[19] I. Masaaki, Z. Lin, L. Tadashi, F. Seiji, J. Jpn. Inst. Met. 67 (2003) 456.
[20] D.M. Bromley, Hydrogen Embrittlement Testing of Austenitic Stainless Steels
SUS 316 and 316L, The University of British Columbia, Vancouver, 2008.
[21] J.K. Tien, S.V. Nair, R.R. Jensen., in: I.M. Bernstein, A.W. Thompson (Eds.),
Metallurgical Society of AIME, Warrendale, Pa, 1980, p. 37.
[22] J. Talonen, P. Aspegren, H. Hnninen, Mater. Sci. Technol. 20 (2004) 1506.
[23] H.K. Birnbaum, P. Sofronis, Mater. Sci. Eng. A 176 (1994) 191.
[24] P. Sofronis, Y. Liang, N. Aravas, Eur. J. Mech. A/Solids 20 (2001) 857.
[25] I.M. Robertson, Eng. Fract. Mech. 68 (2001) 671.
[26] T. Michler, J. Naumann, Int. J. Hydrogen Energy 35 (2010) 821.
[27] T. Kanezaki, C. Narazaki, Y. Mine, S. Matsuoka, Y. Murakami, Int. J. Hydrogen
Energy 33 (2008) 2604.
[28] LW Tsay, Young MC, Chen C, Corros. Sci. 45 (2003) 1985.
[29] L.W. Tsay, J.J. Chen, J.C. Huang, Corros. Sci. 50 (2008) 2973.
[30] L.W. Tsay, Y.C. Liu, M.C. Young, D.Y. Lin, Mater. Sci. Eng. A 374 (2004) 204.
[31] V.N. Shivanyuk, J. Foct, V.G. Gavriljuk, Scr. Mater. 49 (2003) 601.
[32] S.M. Teus, V.N. Shivanyuk, V.G. Gavriljuk, Mater. Sci. Eng. A 497 (2008) 290.
[33] C.M. Ransom, P.J. Ficalora, Metall. Trans. A 11 (1980) 1980.
10 m 10 m
10 m
10 m
20 m
10 m
Fig. 10. Fractographs of various SS specimens: (a) 304L-25a; (b) 304L-25h; (c) 310S-25h; (d) 410A-25h; (e) 410T-25h and (f) 304L-80h.
C.L. Lai et al. / Materials Science & Engineering A 584 (2013) 1420 20

También podría gustarte