Sei sulla pagina 1di 12

Corrosion Science 77 (2013) 210221

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Effect of low temperature on hydrogen-assisted crack propagation


in 304L/308L austenitic stainless steel fusion welds
H.F. Jackson , C. San Marchi, D.K. Balch, B.P. Somerday
Sandia National Laboratories, Livermore, CA 94550, USA

a r t i c l e

i n f o

Article history:
Received 5 February 2013
Accepted 5 August 2013
Available online 11 August 2013
Keywords:
A. Stainless steel
B. SEM
C. Hydrogen embrittlement

a b s t r a c t
Effects of low temperature on hydrogen-assisted cracking in 304L/308L austenitic stainless steel welds
were investigated using elasticplastic fracture mechanics methods. Thermally precharged hydrogen
(140 wppm) decreased fracture toughness and altered fracture mechanisms at 293 and 223 K relative
to hydrogen-free welds. At 293 K, hydrogen increased planar deformation in austenite, and microcracking
of d-ferrite governed crack paths. At 223 K, low temperature enabled hydrogen to exacerbate localized
deformation, and microvoid formation, at austenite deformation band intersections near phase boundaries, dominated damage initiation; microcracking of ferrite did not contribute to crack growth.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Austenitic stainless steels are used extensively in the nuclear
power, chemical processing, and oil and gas industries, in applications where they are susceptible to stress corrosion cracking (SCC)
[13]. SCC begins with an electrochemical process, such as local
anodic dissolution, however synergistic processes of anodic dissolution and cathodic hydrogen damage are thought to contribute to
SCC in stainless steels [4,5]. Hydrogen damage is typically characterized by a signicant loss of tensile ductility. In various alloy/
solution systems, interactions between dislocations at a crack tip
and corrosion products, such as hydrogen, can play a key role in
SCC. Detailed fractographic and microstructural studies support
the premise that adsorbed or absorbed hydrogen promotes crack
growth by localized plastic ow [3,610]. This suggests that effects
of a corrosive environment and sustained tensile stress are made
worse by the presence of hydrogen. Understanding the contribution of hydrogen to deformation and fracture processes is therefore
central to developing a mechanistic understanding of SCC.
Austenitic stainless steels are highly resistant to embrittlement
promoted either by hydrogen or low temperature, retaining significant ductility and fracture toughness in these environments
[4,1115]. Austenitic stainless steel welds can be more susceptible
than base materials to fracture, corrosion, and other degradation.
The compositions of austenitic welds are typically selected to promote primary solidication as d-ferrite and solid-state transformation to austenite to suppress solidication cracking [16]. The as Corresponding author. Present address: Structural Integrity Associates, Inc., San
Jose, CA 95138, USA. Tel.: +1 408 833 7201.
E-mail address: hjackson@structint.com (H.F. Jackson).
0010-938X/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2013.08.004

welded microstructure usually retains several volume percent of


the bcc ferrite phase at room temperature, resulting in a duplex
austenitic/ferritic microstructure. Studies of duplex stainless steels
emphasize their attractive mechanical and corrosion properties,
intermediate between those of austenite and ferrite. High hydrogen fugacities at the metal surface promote hydrogen pickup, especially in applications with cathodic protection or sour gas or oil
containing signicant H2S [1,2]. For hydrogen-exposed austenitic
welds and other duplex (two-phase) austenite/ferrite microstructures, the detrimental inuence of ferrite on room-temperature
ductility and fracture toughness has been noted in several studies
[1723]. Hydrogen transport is complicated by the lower solubility
but higher diffusivity and permeability of hydrogen in the ferrite
phase [24]. Austenitic steels are not as readily embrittled by hydrogen as ferritic steels [2]. Low temperature can induce a ductile
brittle transition in ferrite and localizes deformation in austenite,
yet few studies are reported for hydrogen-assisted fracture of
austenitic stainless steel welds at low temperature, particularly
studies that quantify crack growth resistance by fracture mechanics methods. Mills [25] reviewed cryogenic fracture toughness of
AISI 304 and 316 base metals and their welds at temperatures as
low as 4 K, however, studies at less severe sub-ambient temperatures are needed.
The mechanism of environment-assisted crack propagation is
known to differ depending on the specic combination of metallurgical, environmental, and mechanical variables [5,26]. For example,
features of hydrogen damage in austenitic and duplex stainless
steels differ when they result from exposure to high pressure
hydrogen gas versus cathodic charging in an aqueous solution; or
upon tensile straining after precharging with hydrogen versus in
situ straining in a hydrogen environment [1,2730]. Austenitic al-

211

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

loys susceptibility to embrittlement has been correlated with a


propensity for localized deformation, which can arise due to metallurgical factors such as low stacking fault energy (SFE) [3133] or
environmental factors such as low temperature [3337] or hydrogen exposure [4,38,39]. Hydrogen-assisted crack growth is sensitive to the combination of metallurgical, environmental, and
mechanical variables, hence it is important to understand effects
of relevant combinations of material and environmental variables
(absorbed hydrogen, low temperature, ferrite/austenite phase distribution) on fracture of austenitic welds. In the present manuscript, we have isolated effects of a high-purity gaseous hydrogen
environment. However, hydrogen derived from aqueous sources
(e.g. corrosion reactions at a stress-corrosion crack tip) is thought
to affect alloys in a similar manner [4,14,38].
The present study expands on the results of a previous investigation of room-temperature hydrogen-assisted crack propagation
in gas-tungsten arc (GTA) welds having a 304L base metal and
308L ller [21]. The objectives of the present study are to characterize hydrogen-assisted crack propagation of these welds at low
temperature. Fracture mechanics specimens were thermally precharged in hydrogen gas, and fracture initiation toughness and
crack growth resistance curves were measured at 223 K (50 C).
The effects of low temperature on hydrogen-assisted crack initiation and propagation mechanisms were assessed via electron
microscopy of fracture surfaces and fracture proles.
2. Experimental
The processes of weld fabrication, hydrogen precharging, and
fracture testing of welds at room temperature were detailed by
Jackson et al. [21] and are summarized here.
2.1. 304L/308L Welds
The weld, referred to as 304L/308L throughout this work, was
fabricated from a round bar (64 mm diameter) of annealed 304L
stainless steel having a longitudinal U-shaped groove. The groove
was lled with 1.1 mm diameter 308L ller wire using 9 gas-tungsten arc (GTA) weld passes. Bulk elemental compositions of the
bar, wire, and weld fusion zone are reported in Table 1. The volume
fraction of ferrite in the weld fusion zone was estimated by magnetic measurement (Feritscope, Helmut-Fischer GmbH, Sindelngen, Germany) at several distances from the weld root along the
centerline. The as-welded microstructure was imaged using optical
and scanning electron microscopy (SEM, JSM-840, JEOL, Tokyo, Japan). The elemental compositions of the ferrite and austenite
phases within the weld were quantied by electron probe microanalysis (EPMA, JXA-8200, JEOL, Tokyo, Japan). Variation in tensile
properties as a function of depth in the 304L/308L weld was characterized by tensile tests of subsized specimens which were extracted from the root, mid-height, and top of the weld. Upon
going from the root to the top of the weld, yield strength decreases

from 420 MPa to 350 MPa, while strain at failure increases from
40% to 50% [21].
2.2. Preparation of fracture specimens
Disk-shaped compact-tension (CT) specimens, illustrated schematically relative to the weld in Fig. 1, had a width W (distance
from the load-line to the back face of the specimen) of 40.6 mm,
thickness B of 6.4 mm, and net thickness Bn (between the sidegrooves) of 4.8 mm, consistent with ASTM E1820 [40] and ASTM
E1737 [41]. Specimens tested at room temperature in the noncharged condition had the same W but with a B of 12.7 mm and
Bn of 9.9 mm.
In order to locate the fatigue precrack tip in a similar microstructure in all specimens, the precrack was grown to the same distance into the weld fusion zone, about 1.5 mm. To accommodate
variation in the distance from the notch tip to the root of the weld
in the machined specimens, the nal crack length-to-width ratio
varied between 0.630.68. Precracks were grown in air along the
radial direction of the welded bar along the weld centerline at
10 Hz under a load ratio of 0.1 and nal maximum stress intensity
factor Kmax of 30 MPa m1/2.
Residual stresses in precracked specimens are minimized by
conducting precracking in accordance with standard procedures
in ASTM E1820 [ref]. Specimen dimensions conformed to allowable
dimensions (with respect to specimen geometry) of the starter
notch, nal precrack, and amount of fatigue crack growth. The Kmax
applied during precracking was kept well below the material fracture toughness that is measured during subsequent fracture testing. During the precracking procedure, the displacement is kept
constant, hence the Kmax and DK range decrease progressively as
the crack grows.
Thermal precharging of CT specimens was in 99.9999% hydrogen gas at 138 MPa and 573 K for a minimum of 39 days and up
to 61 days. The charging time and temperature were selected to
achieve at the specimen midthickness a minimum hydrogen concentration of 90% of the equilibrium hydrogen concentration at
the surface. After hydrogen precharging but before mechanical
testing, specimens were stored at below 250 K to minimize hydrogen outgassing. The specimen was equilibrated at the test temperature, either 293 or 223 K, for at least 30 min prior to mechanical
testing. The hydrogen concentration in the fusion zone was
140 wppm (0.8 at.%), as measured by inert gas fusion (Wah
Chang, Albany, OR). This concentration is consistent with that predicted based on the thermal precharging parameters and the
hydrogen solubility of 300-series alloys [42].
2.3. Fracture mechanics testing
Elasticplastic fracture mechanics tests of the fatigue-precracked disk CT specimens were conducted according to ASTM
E1820 [40]. Three or four replicate specimens were tested for each

Table 1
Elemental compositions (wt%) of base metal, ller, and weld fusion zone (balance Fe).

304L
308La
304L/308L weld b,c
304L/308L weldb,d
a
b
c
d
e

Cr

Ni

Mn

Si

Mo

Cu

Nb

Creqe

Nieqe

19.85
20.5
21.27
20.75

10.73
10.3
10.23
10.19

1.6
1.56
1.69
1.65

0.57
0.5
0.51
0.51

0.17
<0.01
0.05
0.04

0.030
0.028
0.04
0.02

0.001
0.012
0.02
0.011

0.005
0.006
0.006
0.005

0.02
0.055
0.047
0.048

0.018
0.02
0.02

<0.005
0.01

19.85
20.51

12.18
12.38

Per manufacturer certication.


By emission spectroscopy.
Analysis at mid-height of weld.
Analysis at root of weld
Nieq = Ni + 35C + 20N + 0.25Cu, and Creq = Cr + Mo + 0.7Nb per Kotecki and Siewert [43].

212

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

tiation of crack extension following blunting was identied as the


second slope change in the DCPD vs. COD record, which is consistent with observations from interrupted fracture mechanics tests
of stainless steels [22]. The fracture initiation toughness was identied from the intersection of the 0.2 mm offset blunting line with
the R curve.
2.4. Post-testing characterization
Fracture surfaces and fracture proles parallel to and normal to
the crack growth direction were characterized by SEM. Elemental
analysis of compositional partitioning by energy dispersive X-ray
spectroscopy (EDS) was used to infer phases in the weld fusion
zone.
Fig. 1. Optical image of the 304L/308L weld in cross-section, with the orientation
and location of disk-shaped CT specimens illustrated schematically.

3. Results
3.1. As-welded microstructure

condition. Room-temperature tests were conducted in the ambient


laboratory air. Tests at 223 K were conducted within an environmental chamber mounted in the load train of a servo-hydraulic test
frame (MTS Corporation, Eden Prairie, MN). Evaporated liquid
nitrogen was used as a coolant, with a resistive heater and temperature controller maintaining the test temperature of 223 K.
Testing was conducted at a constant (monotonic) displacement
rate of 0.4 mm min1. The crack-opening displacement (COD) at
the load-line was measured with a clip gauge, and crack extension
was monitored continuously using the direct current potential
drop (DCPD) method of ASTM E1737 [41]. The test was terminated
manually after a greater than 10% load drop from peak load and an
increase in DCPD by at least 15% from its initial value, or when
load-line displacement exceeded the measurement capacity of
the clip gauge (45 mm).
Following the test, CT specimens were heat-tinted at 623 K for
60 min to mark the extent of crack growth during the test. The
specimens were then broken apart in liquid nitrogen, and the nal
crack length and precrack length were measured from the fracture
surfaces. The calculated crack length based on DCPD data was linearly scaled so that the nal calculated and measured crack lengths
were equal. Prior to correction, the discrepancy between the nal
measured and calculated crack lengths was less than 1.5%.
The J-integral crack growth resistance (JR) curve was constructed from the load vs. COD and DCPD vs. COD records. The initial J vs. crack growth increment (Da) relationship was assumed to
be a linear crack blunting response following ASTM E1737. The ini-

The microstructure of the fusion zone consisted of primary dferrite in a matrix of austenite (c), as seen in Fig. 2. The ferrite
number (FN) as measured at distances of 2, 6, 10, and 14 mm from
the weld root along the centerline was 9.4, 10.1, 10.7, and 10.1,
respectively. These values represent the mean of ve measurements per location. The FN approximates the volume percent ferrite, and the measurements are consistent with values of 810
predicted from the WRC-1992 diagram [43], based on the calculated Creq and Nieq for the weld metal (Table 1).
Table 2 shows the elemental compositions of the austenite and
ferrite phases within the weld, as measured by EPMA spot analyses
at two locations: at mid-height of the weld and near the weld root.
Table 2 also compares the compositions of the individual phases to
the average composition of the weld fusion zone, as measured by
emission spectroscopy in the same locations.
In the fusion zone, the average composition by emission spectroscopy closely matched that of the original 308L ller wire. Weld
austenite had approximately the same wt% Ni and Cr as in the 304L
and 308L starting materials, while the approximately 810 vol% of
d-ferrite was depleted in Ni and enriched in Cr.
3.2. Fracture toughness
Compared to the non-charged condition, precharged hydrogen
markedly decreased the fracture initiation toughness of 304L/
308L welds at both 293 and 223 K, as shown by crack growth resis-

Fig. 2. (a) Optical and (b) backscattered electron images of the fusion zone in the 304L/308L weld near the fusion zone/base metal interface at the weld centerline. The
solidication direction is nominally from bottom to top of the image. The microstructure contains skeletal ferrite (d) in an austenite (c) matrix.

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221


Table 2
Elemental compositions (wt%) of ferrite and austenite phases within the weld.
Phase

a
b
c

Location

Cr

Ni

Weld ferrite

Weld mid-height
Weld roota

28.17
27.86

4.41
4.38

Weld austenite

Weld mid-heighta
Weld roota

20.02
19.73

10.24
10.40

Weld average

Weld mid-heightb
Weld rootb
308L ller wirec

21.27
20.75
20.5

10.23
10.19
10.3

Base metal

Outside fusion zonea


304L annealed barc

18.44
19.85

10.21
10.73

EPMA, mean of three measurements for ferrite and nine for austenite.
Emission spectroscopy.
Per manufacturer certication.

tance (JR) curves (Fig. 3) and fracture initiation toughness (Jmax or


JIH) values (Table 3). Also reported in Table 3 are the corresponding
stress-intensity factor KJ values, determined by:

KJ

r
EJ
1  m2

where E is Youngs modulus, J is fracture initiation toughness, and m


is Poissons ratio [40]. The JR curve slope over the rst 0.5 mm of
crack extension immediately following blunting is reported as crack
growth resistance dJ/dDa.

213

Welds not exposed to hydrogen and tested at both 293 and


223 K exhibited signicant crack tip blunting, and tests were terminated manually before the onset of noticeable crack extension,
as conrmed by post-test analysis. Crack extension in non-charged
specimens was insufcient to permit calculation of JIC, and instead
we report as Jmax the lesser of two values, either (1) the maximum J
value recorded during the test or (2) the maximum J-integral
capacity of the specimen as permitted by ASTM E1820 [40], i.e.
the lesser of b0rY/10 or BrY/10, where b0 is the original uncracked
ligament, B is the original specimen thickness, and rY is the material ow stress. These Jmax values represent lower-bound fracture
initiation toughness measurements.
It is important to note that the difference in Jmax between noncharged welds tested at 293 and 223 K (Table 3) was primarily due
to the difference in specimen thickness and the corresponding
dependence of the maximum J-integral capacity on this dimension.
Non-charged specimens tested at 223 K were thinner than those
tested at 293 K, while hydrogen-precharged specimens were all
of the same thickness. Regardless, crack blunting and a lack of
noticeable crack extension were observed for non-charged specimens tested at both 293 and 223 K.
The measured fracture initiation toughness of hydrogen-precharged welds represents a reduction of greater than 59% compared to the lower-bound fracture initiation toughness of noncharged welds. Hydrogen-precharged specimens met the dimensional requirements in ASTM E1820 [40] to dene conditional JQ
as JIC. To reect the hydrogen-precharged condition, fracture initiation toughness is denoted JIH and stress-intensity factor is KJIH.
Fracture initiation toughness with internal hydrogen is essentially the same at 293 and 223 K. At the initiation of crack growth
(Fig. 3b), the JR curves for both temperatures are indistinguishable. However, at longer crack extensions (Fig. 3a), welds tested
at the lower temperature exhibit lower crack growth resistance
dJ/dDa. This divergence in JR curve slope occurs at a crack extension of about 0.81 mm from the precrack tip. This distance was
found to correspond to a change in orientation and morphology
of austenite grains and d-ferrite dendrites at a depth in the fusion
zone where two weld passes overlapped (Fig. 4). Crack paths were
tortuous in general, and any crack deection or differences in fracture appearance due to microstructural differences in the weld
pass overlap region were not visibly distinguished from that in
other regions.
3.3. Fractography

Fig. 3. Crack growth resistance (JR) curves for 304L/308L GTA welds with and
without precharged hydrogen, tested at 293 K and 223 K. Up to four JR curves per
condition are plotted. (a) At longer crack extensions, welds tested at 293 K have
greater crack growth resistance than at 223 K. (b) Immediately following crack
initiation, JR curves were indistinguishable at 293 versus 223 K, as were JIH values.

Fracture features are signicantly different for specimens fractured with and without hydrogen; for specimens fractured in
hydrogen, a different distribution of fracture features is seen when
test temperature is decreased from 293 to 223 K. Effects of hydrogen and test temperature on fracture features are compared in
Fig. 5 and in higher magnication in Fig. 6.
Non-charged welds at both 293 and 223 K exhibited signicant
crack blunting followed by a slight amount of ductile crack extension. These fracture surfaces featured only equiaxed and U-shaped
dimples (Fig. 5a and b), indicating microvoid coalescence (MVC).
Under these conditions, microvoids nucleate at ne spherical
precipitates.
Hydrogen exposure altered the fracture mode of austenitic
welds (Fig. 5c and d). In hydrogen-precharged welds fractured at
293 K, the fracture surface was dominated by features which were
elongated and aligned parallel to the crack propagation direction,
which was also the weld solidication direction (Figs. 5c, 6a and
b). The dendritic morphology of these features (denoted by D in
Figs. 5c and 6a) resembled the underlying d-ferrite microstructure
(Fig. 2), with smooth, relatively at features interspersed with tear
ridges (denoted by T in Fig. 6a), i.e. vertical steps joining two at

214

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

Table 3
Fracture initiation and crack growth toughness of 304L/308L GTA welds
Non-chargeda,b

Hydrogen-precharged
2

a
b
c

0.5

JIH (kJ m )

KJIH (MPa m

223 K

144
123
109

175
161
152

171
175
181

Mean

125

163

176

293 K

110
123
132
111

153
161
167
154

209
190
196
242

Mean

119

159

209

dJ/dDa

1

(kJ m mm

2

Jmax (kJ m

dJ/dAac (kJ m2 mm-1)


0.5

KJ (MPa m

dJ/dDa

304

254

965

625

364

965

(kJ m2 mm1)

Mean of 3 replicate tests 6 of non-charged specime.


B of non-charged specimens was 6.35 mm for tests at 50 C and 12.7 mm for tests at 20 C.
Slope of JR curve, measured over 0.5 mm of crack extension immediately following blunting.

Fig. 4. A change in microstructural orientation and morphology was observed where two weld passes overlap, about 0.81 mm from the precrack tip and 2.32.5 mm from
the base metal/weld interface. Backscattered electron image.

Fig. 5. Fracture surfaces of hydrogen precharged and non-charged welds, tested at 223 or 293 K. Without hydrogen, at (a) 293 K or (b) 223 K, fracture surfaces had only
dimpled rupture. With hydrogen at (c) 293 K, the fracture surface was dominated by elongated features with a dendritic morphology (D) and aligned with the crack growth
direction, while at (d) 223 K, the dominant features were elongated (E) and equiaxed (Q) dimples, with few dendritic-shaped features. Crack growth direction is from bottom
to top in all images.

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

215

Fig. 6. Fracture features at matching locations on both halves of hydrogen-precharged fracture surfaces. At 293 K (a and b), the dominant features were dendritic (D) with
tear ridges (T) between dendritic features. At 223 K (c and d), few regions had dendritic features (D). The dominant features (e and f) were planar arrays of elongated (E)
dimples and patches of equiaxed (Q) dimples. Crack growth direction is from bottom to top in all images.

areas of dimpled rupture, with steps also exhibiting dimples. Most


of the fracture surface exhibited features related to the underlying
dendritic structure of ferrite and the inter-dendritic regions of
austenite.
In hydrogen-precharged welds fractured at 223 K (Figs. 5d, 6c,
and d), considerably fewer regions of the fracture surface showed
evidence of the dendritic microstructure (denoted by D in Fig. 6c
and d). Most regions showed either elongated or equiaxed dimples
(denoted by E and Q respectively in Figs. 5d and 6e). Elongated
dimples occurred in nominally parallel and coplanar arrays
(Fig. 6e and f). Parallel arrays of elongated dimples formed steps
on the fracture surface (Figs. 5d and 6e).
Matching locations on both fracture surfaces of hydrogen-exposed welds (Fig. 6) were analyzed by EDS to provide a basis for
inferring the phase that was present at the surface based on the
approximate Ni and Cr concentrations, which were consistent with
the values for austenite or ferrite in Table 2. EDS analysis alone
may be inconclusive if the ferrite dimensions are very thin and ferrite evades detection by the EDS probe. By comparing elemental
analyses and fracture features on both halves of the fracture surface, the crack path could be identied as fracture within ferrite,
at the austenite/ferrite phase boundary, or within austenite. All

three types of crack path were observed in various locations at


both temperatures.
In the present work, no distinction was made between cracks at
c/c boundaries and cracks within a c grain, although both may occur. For this reason, it cannot be conclusively stated that d was absent at a given fracture location and that the crack path was
entirely contained within the c phase.
Planar arrays of elongated dimples, seen only at 223 K, were
associated with fracture within the austenite phase; such crack
growth could be either transgranular or at c/c interfaces. Previously [20,21], it was established that for hydrogen-precharged
austenitic welds at 293 K, tear ridges and equiaxed dimples were
associated with fracture within austenite; the at areas containing
dendritic features were associated either with fracture at the austenite/ferrite phase boundary or within ferrite.
At both temperatures, fracture surfaces exhibited signicant
out-of-plane crack growth and tortuous crack topographies. With
hydrogen at both 293 and 223 K, crack extension evolved as multiple discontinuous microcracks that joined up to form the main
crack path. At 293 K, the microcracks formed on planes parallel
to the Mode I crack growth plane, i.e. normal to the tensile axis
[21]. At 223 K, microcracks formed primarily on planes inclined

216

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

to the Mode I crack growth plane, as conrmed from stereo pair


fractographs.
3.4. Fracture prole metallography of hydrogen-exposed welds
Fracture proles parallel and transverse to the crack growth
direction in hydrogen-exposed welds were imaged by backscattered electron microscopy (Figs. 7 and 8). Atomic number contrast
distinguishes d-ferrite as the dark grey dendritic phase and austenite as the light grey matrix. Localized deformation bands appear as
bright parallel streaks in the light grey austenite matrix.
Fracture proles (Fig. 7) suggested that, although microvoids/
microcracks initiated near ferrite at both 223 and 293 K in hydrogen, crack growth occurred by different mechanisms at the two
temperatures.
At 293 K, microcracks were predominantly associated with dferrite (Fig. 7a and b). Microcracks nucleated either within the ferrite phase or at the austenite/ferrite phase boundary (Fig. 7a) and
propagated parallel to ferrite dendrites (denoted by d in Fig. 8a).
These microcracks often created coarse cracks (Fig. 7b) parallel to
the main crack, which were located up to several hundred micrometers above or below the fracture surface (Fig. 8c).
At 223 K, there was no evidence for large microcracks away
from the fracture surface. Fine microvoids 1 lm or less in diameter
(Fig. 7d and f) were the main type of microstructural damage ob-

served. With an equiaxed cross-section, the three-dimensional


shape of these microvoids could be either spherical or tubular.
Some ne porosity was randomly distributed in the as-welded
microstructure, however the microvoids that were associated with
damage accumulation and fracture had a distinctive arrangement.
Specically, these microvoids tended to be aligned and to coincide
with the intersections between deformation bands (Fig. 7f). The
predominant microvoid nucleation site was adjacent to d-ferrite
(Fig. 7cf), while a smaller fraction formed elsewhere in austenite
grains. Microcracks evolved via the nucleation and coalescence of
closely-spaced microvoids (Fig. 7c, d, and f).
Fracture proles reveal a second key difference between fracture at 293 K and at 223 K: the macroscopic crack path. At both
293 and 223 K, fractures had signicant out-of-plane crack growth
and tortuous crack topographies (Fig. 5c and d). This is reected in
cross-sections as macroscopic steps (denoted by S in Fig. 8a,c,d).
At 293 K, steps corresponded to microcracks growing along dferrite on various planes ahead of the crack tip that were nominally
parallel to each other and perpendicular to the tensile axis (Fig. 8a
and c). The main crack grew when these microcracks joined along
planes parallel to the tensile axis, facilitated by intense shear in the
ligaments between them (denoted by I in Fig. 8a). This out-of-plane
crack growth corresponds to that seen on fracture surfaces
(Fig. 5c).

Fig. 7. Fracture proles show different mechanisms of microvoid nucleation at 293 versus 223 K. At 293 K (a and b), microcracks originate at and propagate along d-ferrite. At
223 K (cf), microvoids nucleate where localized deformation bands intersect other deformation bands, and microvoids are concentrated near phase boundaries. Microcracks
form via the coalescence of multiple microvoids (c, d, and f). Crack growth direction is from left to right in all images except for a, in which crack growth is normal to the page.

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

217

Fig. 8. Fracture proles reveal that at both temperatures, the crack path is marked by macroscopic steps (S). At 293 K (a and c), cracks propagate along ferrite dendrites (d) but
can link up through intense shear (I) in the ligaments between them. At 223 K (b and d), crack paths are primarily within austenite and do not clearly follow ferrite.

At 223 K, macroscopic steps and out-of-plane crack growth


were also observed. However, microcracks were observed to form
on planes inclined to the tensile axis (Fig. 6cf). In cross-section,
the crack path does not clearly follow the ferrite phase (Fig. 8b
and d) and is primarily within the austenite phase.

4. Discussion
As shown in previous studies [20,21], one primary role of
hydrogen in degrading fracture resistance and altering fracture
mechanisms in austenitic stainless steel welds at room temperature is enhancing planar deformation in austenite. This phenomenon produces stress concentrations where localized
deformation bands intersect d-ferrite. These stress concentrations
facilitate microcracking of the elongated d-ferrite grains, which
are aligned with the direction of crack growth. This dominant
role of ferrite in the initiation and propagation of microcracks
in austenitic welds at room temperature is illustrated
in Figs. 6a, 6b, 7a, andb and documented in several studies
[1721].
Extending the study on 304L/308L welds at room temperature
[21], we found that low-temperature (223 K) exposure had no effect on the fracture initiation toughness of hydrogen-precharged
welds (Table 3), although the fracture appearance was clearly
changed (Fig. 5c and d). In contrast to room-temperature fracture,
at 223 K fracture surfaces and fracture proles showed a diminished role of d-ferrite in facilitating crack propagation (Figs. 6
and 7), relative to room temperature.
We propose that low temperature alters the competition between potential fracture initiation modes in hydrogen-precharged
austenitic welds. The two main effects of low temperature are that
(1) low temperature localizes plastic deformation in austenite, rendering this phase more susceptible to hydrogen-enhanced localized plasticity, and (2) low temperature increases the hydrogen
concentration in austenite near phase boundaries through its effects on the thermodynamics of hydrogen trapping. In the following sections, we discuss effects of low temperatures on plastic

deformation and hydrogen distribution in welds and how these


changes govern the fracture mode.
4.1. Sequence of crack initiation and propagation in hydrogenprecharged welds at 223 K
In the presence of hydrogen, low temperature (223 K) produced
a signicantly different fracture surface than seen at room temperature (293 K).
At 293 K, a mixture of ferrite cleavage and phase boundary
separation was observed. Microcracks nucleated and propagated
preferentially at d-ferrite (Figs. 7a and b), resulting in elongated,
dendritic features on fracture surfaces (Figs. 5c, 6a, and 6b)
that resemble the underlying ferrite morphology (Fig. 2). These
microcracks formed in the ferrite phase itself or at austenite/
ferrite interfaces, facilitated by high stress concentrations at these
sites.
At 223 K, neither ferrite cleavage nor phase boundary separation is an important fracture mechanism, based on the observed
fracture surface features (Fig. 6cf). Rather, the dimpled surface
indicates that fracture was caused by MVC within the austenite
phase. Microcracks originated as microvoids at intersections between localized deformation bands in austenite (Fig. 7cf). Because
deformation bands are parallel, intersections between deformation
bands occur in rows, as do the microvoids that form at these intersections. The non-equiaxed growth of these aligned microvoids
produced coplanar arrays of elongated fracture dimples (Figs. 5d,
6e, and 6f). In limited locations, the fracture appearance at 223 K
resembled the dendritic features associated with microcracks in
d-ferrite or at austenite/ferrite boundaries at 293 K (denoted by D
in Fig. 6c and d).
An important feature of the proposed low-temperature fracture mechanism is that, although deformation band intersections occurred throughout the austenite bulk, microvoids were
not randomly distributed but were concentrated near austenite/ferrite phase boundaries (Fig. 7cf). We can infer that in
welds or two-phase microstructures, two conditions promote
microvoid nucleation in austenite: (1) deformation bands inter-

218

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

sect and (2) the intersection is in close proximity to an austenite/ferrite phase boundary. The role of hydrogen in facilitating
void formation near phase boundaries will be discussed in the
nal section.
Despite the differences in fracture appearance, the fracture process at 223 K had a similar sequence of steps compared to the fracture process at 293 K [20,21].
First, planar deformation in austenite developed as localized
bands. The formation of these deformation bands is facilitated by
the hydrogen enhanced localized plasticity (HELP) mechanism
[4446]. In austenite, HELP results in greater slip planarity
[4,38,39], exacerbating the effects of metallurgical factors predisposing an alloy to planar slip, such as low stacking fault energy
(SFE) [3133]. In metastable 300-series alloys such as 304L and
308L, SFE decreases as nickel content decreases [47,48]. It follows
that the effects of low SFE, low temperature, and hydrogen would
produce more severely localized slip than any factor acting independently. As slip becomes more localized, other planar deformation modes such as twinning and e-martensite become activated
at lower strains [33,35,4953].
The next step in the process was the impingement of localized
deformation bands on microstructural obstacles, thus producing
stress concentrations. Austenitic stainless steels contain various
microstructural features that can act as obstacles to slip or twinning, including grain boundaries, secondary phases, and annealing
twin boundaries. At 293 K, the d-ferrite phase could serve as a barrier to propagation of planar deformation bands, and the resulting
stress concentration facilitated microcrack initiation either as
cleavage of ferrite or decohesion at the austenite/ferrite phase
boundary [21]. At 223 K, the resulting microstructural damage
was instead a series of discrete microvoids at deformation band
intersections near phase boundaries. Hence, low temperature
changed the dominant mode of damage accumulation, leading to
a different mode of microcrack initiation relative to that at room
temperature.
The nal step was crack propagation via the growth and coalescence of microcracks. At 293 K, microcracks formed from cleavage
in ferrite or decohesion of austenite/ferrite interfaces, then these
microcracks linked by intense shear in the remaining ligaments
(Figs. 7a, 7b, 8a, and 8c). At 223 K, microvoids nucleated predominantly adjacent to weld ferrite, while ferrite cleavage and austenite/ferrite interface decohesion were infrequently observed.
Instead, under low-temperature conditions, decohesion was not a
dominant fracture mode; an alternative mode of crack propagation
intervened, and multiple microvoid nucleation and coalescence
events were required for microcracks to form (Fig. 7cf). Similar
to the process at room temperature, individual microcracks ahead
of the crack tip at low temperature linked up via shear in the ligaments between them, producing coarse steps on the fracture surfaces (Fig. 8b and d).
4.2. Effect of temperature on crack initiation: localized deformation
In this study, temperature governed the mode by which hydrogen-assisted cracks initiated and propagated. At both 293 and
223 K, planar deformation and microstructural obstacles played a
prominent role in the initiation of hydrogen-assisted cracking in
austenitic stainless steel welds.
We propose that low temperature altered fracture mechanisms in hydrogen-precharged welds in part by conditioning the
austenite for more severe planar slip that is further exacerbated
by hydrogen. The planar deformation bands that evolve increase
the severity of stress concentrations at intersecting deformation
bands near the austenite/ferrite phase boundary, perhaps elevating stress concentrations at these sites relative to the stress concentrations at intersections of deformation bands with d-ferrite.

Consequently, void nucleation in austenite near phase boundaries


is favored over microcracking in ferrite or at austenite/ferrite
interfaces.
Low temperature exacerbates localized deformation in several
ways. First, low temperature increases the tendency toward planar
slip in austenite by decreasing SFE [3337]. More intense planar
slip on [1] planes in fcc austenite gives rise to deformation twins,
bcc a0 -martensite, and hcp e-martensite at lower strains
[33,35,4953]. These deformation modes promote the connement
of deformation in discrete bands, leading to more pronounced
strain incompatibility and enhanced stress concentration at obstacles. After deforming 301LN stainless steels in tension at temperatures between 233 and 353 K, Talonen and Hanninen [33] observed
that, as temperature decreased, a greater volume fraction of a0 martensite formed, and at lower levels of strain. They proposed
that low SFE at low temperature facilitates the martensitic transformation. In contrast, at higher temperature, the overlapping of
stacking faults is less regular, hindering the formation of martensite nuclei. Deformation at lower temperature produces more twins
and martensite the results of planar deformation modes than at
room temperature.
Second, low temperature could also change the types of deformation products that preferentially form. In addition to decreasing
SFE, low temperature could alter the free-energy differences between the fcc, bcc, and hcp phases, and thus the thermodynamic
driving force for transformation to the hcp or bcc martensite
phases. During an investigation of tensile deformation of austenitic
steels containing 16-18 wt% Cr and 1113 wt% Ni at temperatures
between 77 and 523 K, Lecroisey and Pineau [35] observed deformation bands containing both deformation twins and e-martensite, with a higher proportion of twins at higher deformation
temperatures. They proposed that the tendency to nucleate
austenitic twins versus e-martensite at a given temperature depends on the relative values of extrinsic and intrinsic stacking fault
energies as well as the volume contraction due to the fcc-to-hcp
transformation. Temperature probably determines the relative
proportions of hcp or bcc martensite and fcc twins that compose
deformation bands. Although insufcient data exists to evaluate effects of deformation band makeup on the severity of stress concentrations, it is reasonable to presume that deformation products
such as twins and e-martensite conne deformation in discrete
bands, promoting strain incompatibility and stress concentration
at obstacles [21].
Third, more severe planar deformation has been correlated with
physically thinner deformation twins and e-martensite plates and,
consequently, higher stress concentrations where they impinge on
obstacles.
Based on an investigation of uniaxial tensile fracture of nitrogen-bearing austenitic steels at temperatures between 4 and
298 K, Mullner and coworkers [54,55] attributed fracture initiation
to stress concentrations at deformation twin intersections. Lower
temperature decreased the thickness of twins and the dislocation
density, and they asserted that such conditions rendered twins
effective obstacles against propagation of impinging twins, creating stress concentrations high enough to initiate microcracks at
twin intersections.
Based on fractographs and fracture proles of low-SFE, highMn austenitic steels fractured in Charpy tests at 77 K, Takaki
et al. [56] inferred that intersections between e-martensite
plates and the associated stress concentrations led to the formation of parallel tunnel-like voids at these intersections. They invoked this mechanism to explain the dimpled planes and steplike ridges they observed on fracture surfaces. It is plausible that
propagation of both twins and e-martensite plates could be
blocked by the austenite/ferrite phase boundary as well. Thinner
deformation twins and e-martensite plates produced at lower

219

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

temperatures would lead to more severe stress concentrations


than at room temperature.
While the preceding discussion supports the notion that low
temperature conditions austenitic stainless steels for more localized deformation, which can then be exacerbated by elevated
hydrogen, leading to elevated stress concentrations at intersecting
deformation bands, one detail still requiring attention is the observation that at low temperature, microvoids form predominantly
near austenite/ferrite phase boundaries.
One possible explanation for this observation is that deformation is disproportionately localized near austenite/ferrite phase
boundaries, where Ni depletion locally decreases SFE. Brooks
et al. [18,57] attributed the Ni concentration prole observed
for AISI 309 GTA welds to partitioning during the diffusion-controlled solid-state transformation from ferrite to austenite following primary ferrite solidication. The authors attributed Ni
enrichment at the centers of austenite grains (i.e. the boundaries
of solidication cells) to secondary austenite solidication due to
Ni enrichment of the liquid. EDS concentration proles of the
present weld suggested that Ni concentration was 2.53 wt% lower near the phase boundaries than in the centers of austenite
grains (i.e. the boundaries of solidication cells), where the 12
13 wt% Ni exceeded the nominal Ni concentration in weld austenite, about 10.2 wt% (Table 2). Such concentration gradients could
bias localized deformation toward the austenite/ferrite phase
boundaries.
In summary, low temperature enhances the formation of planar
deformation bands in austenite, especially near austenite/ferrite
phase boundaries. One potential role of hydrogen in linking deformation and fracture is that hydrogen magnies planar slip. This
intensied planar slip or the deformation products that it stimulates (e.g. twinning, e-martensite) leads to stress concentrations
at obstacles such as intersecting deformation bands. The observation that microvoid nucleation in austenite near phase boundaries
out-competes microcrack formation in ferrite could be due in part
to increased stress concentrations at deformation band intersections near phase boundaries relative to those at intersections between deformation bands and d-ferrite. The difference in damage
initiation modes gave rise to the signicant difference in fracture
surface appearance.
While the fracture modes at 293 and 223 K differed signicantly, temperature had no notable effect on fracture initiation
toughness (Table 3). We propose that the similar JIH values at
293 and 223 K are a coincidence. In other words, it is coincidental
that the dominant fracture processes at these two temperatures
are characterized by a similar magnitude of the local fracture criterion, e.g., a critical local strain. For example, at cryogenic temperatures, crack nucleation would likely occur by cleavage of d-ferrite
due to its ductilebrittle transition [25]. Other fracture modes
might dominate at different sub-ambient temperatures, altering
crack initiation toughness.
On the other hand, dJ/da, a measure of crack growth resistance, was slightly lower at 223 K than at 293 K. According to
Ritchie and Thompson [58], metallurgical factors affecting the local fracture strain have a greater effect on crack growth than on
crack initiation, owing to a weaker strain singularity ahead of a
stable growing crack versus a stationary crack. Metallurgical factors dictating ductility include local microstructure and the degree of planar deformation. In the present welds, the JR
curves for tests at 293 and 223 K coincided near initiation but
diverged after cracks extended into a region of overlap between
weld passes and a change in microstructural orientation and
morphology (Fig. 4). This microstructural change and the increase in planar deformation at 223 K had a stronger inuence
on dJ/da at 223 K than at 293 K.

4.3. Effect of temperature on crack initiation: local hydrogen


distribution
We propose that low temperature alters the thermodynamics of
hydrogen trapping in austenitic stainless steel welds in a way that
redistributes hydrogen from the interiors of d-ferrite grains to
phase boundary trap sites. At 223 K, this could contribute to hydrogen-assisted fracture initiation near phase boundaries rather than
within d-ferrite.
The total concentration of dissolved hydrogen consists of
hydrogen in interstitial lattice sites and hydrogen trapped at
microstructural features such as grain boundaries, phase boundaries, dislocations, and impurities [59]. Luppo et al. [60] provided
evidence that austenite/ferrite phase boundaries in welds serve
as hydrogen trap sites. According to Hirth [61], trap sites become
more densely populated as temperature decreases, as expressed
in following equation:

hT
EB
hL exp 
1  hT
kT

where hT is the trap site occupancy, hL is the lattice hydrogen concentration (in this formulation, hL  1), Eb is the binding energy
of hydrogen in the trap site, k is Boltzmanns constant, and T is temperature. Because a hydrogen-precharged specimen can be considered a closed system with a xed total hydrogen concentration, the
hydrogen demanded by trap sites at lower temperature must be
supplied by lattice sites. Conversely, when temperature increases,
trap occupancy decreases, with hydrogen migrating into lattice
sites.
From a kinetic standpoint, the d-ferrite phase is more likely
than austenite to supply hydrogen to trap sites at phase boundaries. Based on the values recommended by Perng and Altstetter
[62] for ferritic stainless steel 29Cr4Mo2Ni and those recommended by San Marchi et al. [42] for 300-series austenitic stainless
steels, the diffusivity of hydrogen in ferrite and austenite were
determined at 293 and 223 K and are summarized in Table 4.
Hydrogen diffusivity is four to six orders of magnitude faster in
bcc ferrite than in fcc austenite at the test temperatures. Even
though the lattice solubility of hydrogen in austenite is about
100 times that in ferrite [63], and thus the concentrations will differ between the two phases, the overall diffusional ux will still be
greater in d-ferrite.
The present analysis demonstrates that hydrogen diffusivity in
d-ferrite is rapid enough to satisfy the increased demand for hydrogen at traps as induced by low temperature. The hydrogen diffusion distance x varies with diffusivity D and time t according to
Eq. (3), derived from the solution to the transient diffusion
equation:

p
x  2 Dt

Specimens were exposed to the test temperature for at least


30 min and for up to several hours prior to mechanical testing.
The hydrogen diffusion distance in d-ferrite during a typical expoTable 4
Hydrogen diffusivity D and diffusion distance x in ferrite and austenite phases at 293
and 223 K
a

(m2 s1)

x (lm)

T (K)

t = 30 min

t=3h

Ferrite

223
293

1  1013
8  1012

30
200

60
600

Austenite

223
293

2  1019
2  1016

0.04
1

0.1
3

a
Based on the values recommended by Perng and Altstetter [62] and San Marchi
et al. [42].

220

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221

sure (Table 4) is signicantly greater than the maximum required


migration distance, i.e. from the center of a ferrite grain to trap
sites at phase boundaries, about 0.5 lm.
At low temperatures, hydrogen migration from lattice sites in dferrite to traps at phase boundaries both depletes hydrogen from
ferrite grains and enriches phase boundaries with hydrogen. This
tilts the competition away from microcrack formation in ferrite
and toward microvoid nucleation near phase boundaries.
As a result of this hydrogen enrichment at phase boundaries
that is induced by low temperature exposure, we propose that
hydrogen has two roles in concentrating damage near phase
boundaries. First, hydrogen in concert with low temperature localizes deformation and increases stress concentrations near phase
boundaries, consistent with the concepts described in the previous
section. Second, hydrogen could have a direct role in facilitating
microvoid nucleation near phase boundaries through several possible mechanisms.
Vacancies are generated by intense localized plastic deformation and stabilized by hydrogen, according to the model of hydrogen-stabilized vacancy formation proposed by Nagumo [64].
Building upon this model, Neeraj et al. [65] propose a nanovoid
coalescence (NVC) micromechanism of hydrogen embrittlement
for ferritic steels. In the NVC mechanism, nanovoids nucleate at a
critical excess vacancy concentration. They proposed that NVC
gave rise to the nanoscale (1020 nm) dimples they observed on
quasi-brittle facets on fracture surfaces.
In the previous section, we linked the formation of microvoids
in austenite to intersections between planar deformation bands
that may consist of e-martensite plates [56] and deformation twins
[54,55]. It is plausible that hydrogen could directly promote fracture by encouraging void formation at these intersections, in addition to hydrogens indirect role in fracture, i.e. localizing
deformation.
In the context of the present work, we propose that hydrogen
enrichment at phase boundaries, particularly at low temperature,
makes near-phase-boundary austenite the preferred site for damage nucleation. At low temperature, this damage proceeds in a
manner consistent with the aforementioned mechanisms, in
which hydrogen stabilizes the vacancies that eventually form
microvoids or hydrogen facilitates microvoid nucleation at the
intersection of localized deformation bands consisting of e-martensite and deformation twins. At room temperature, the damage
is predominantly in the form of microcracks in ferrite or at phase
boundaries.
5. Conclusions
Elasticplastic fracture mechanics tests were conducted to evaluate effects of low temperature on hydrogen-assisted crack propagation in austenitic stainless steel welds fabricated from AISI 304L
base metal and 308L ller metal and which had been thermally
precharged with hydrogen gas:
1. Hydrogen degrades fracture resistance by >59% relative to the
lower-bound fracture initiation toughness of non-charged
welds. At room temperature (293 K), hydrogen alters the dominant mode of damage accumulation in austenitic welds,
thereby altering fracture mechanisms by (1) enhancing localized planar deformation in austenite and (2) facilitating microcracking of d-ferrite or its interfaces. The dendritic
microstructure of weld ferrite, which is aligned with the direction of crack growth, governed the crack path.
2. Low temperature (223 K) exposure altered fracture mechanisms in hydrogen-precharged welds. At 223 K, fracture initiation was dominated by microvoid formation at deformation
band intersections in austenite near phase boundaries, while

microcracking of ferrite and its interfaces did not play a significant role in crack growth. Low temperature was presumed to
(1) lower SFE and condition the austenite for more localized
deformation, which was further exacerbated by hydrogen, and
(2) enrich phase boundaries with hydrogen by increasing the
occupancy of hydrogen at trap sites relative to lattice interstitial
sites in ferrite. The more localized deformation in austenite,
particularly near phase boundaries, could lead to more severe
stress concentrations at deformation band intersections relative
to those at intersections of deformation bands with ferrite.
3. Additionally, we propose that elevated hydrogen concentration
near phase boundaries may directly promote localized void
nucleation in austenite, consistent with the mechanisms of
hydrogen-stabilized vacancies proposed by Nagumo [64] and
vacancy-induced nanovoid coalescence (NVC) proposed by
Neeraj et al. [65]. Alternately, hydrogen facilitated microvoid
formation at intersections between localized deformation
bands, consistent with literature reports of stress concentrations and microvoid nucleation at intersections between e-martensite plates [56] and deformation twins [54,55].
4. Exposure to 223 K had no effect on fracture initiation toughness
of precharged welds, despite the difference in the dominant
mode of damage accumulation. It is likely coincidental that
the dominant fracture modes at the two test temperatures
(microvoid formation at intersecting deformation bands at
223 K vs. microcracking in ferrite or its interfaces at 293 K)
had similar local fracture criteria, resulting in similar fracture
initiation toughness values.

Acknowledgments
The authors are grateful to J. Campbell for hydrogen pressure
systems support, A. Gardea for metallographic preparation, and J.
Chames and R. Nishimoto for SEM imaging. Sandia is a multiprogram laboratory operated by Sandia Corporation, a wholly owned
subsidiary of Lockheed Martin Corporation, for the US Department
of Energys National Nuclear Security Administration under contract DE-AC04-94AL85000.

References
[1] A.A. El-Yazgi, D. Hardie, The embrittlement of a duplex stainless steel by
hydrogen in a variety of environments, Corros. Sci. 38 (1996) 735744.
[2] W. Zheng, D. Hardie, The effect of hydrogen on the fracture of a commercial
duplex stainless steel, Corros. Sci. 32 (1991) 2336.
[3] Y. Mine, T. Kimoto, Hydrogen uptake in austenitic stainless steels by exposure
to gaseous hydrogen and its effect on tensile deformation, Corros. Sci. 53
(2011) 26192629.
[4] A.W. Thompson, I.M. Bernstein, The role of metallurgical variables in
hydrogen-assisted environmental fracture, in: M.G. Fontana, R.W. Staehle
(Eds.), Advances in Corrosion Science and Technology, Plenum Press, New
York, NY, 1980, pp. 53175.
[5] D. Delafosse, T. Magnin, Hydrogen induced plasticity in stress corrosion
cracking of engineering systems, Eng. Frac. Mech. 68 (2001) 693729.
[6] S.P. Lynch, Metallographic contributions to understanding mechanisms of
environmentally assisted cracking, Metallography 23 (1989) 147171.
[7] S.P. Lynch, A fractographic study of hydrogen-assisted cracking and liquidmetal embrittlement in nickel, J. Mater. Sci. 21 (1986) 692704.
[8] S.P. Lynch, A fractographic study of gaseous hydrogen embrittlement and
liquid-metal embrittlement in a tempered-martensitic steel, Acta Metall. 32
(1984) 7990.
[9] Y. Mine, K. Hirashita, M. Matsuda, M. Otsu, K. Takashima, Effect of hydrogen on
tensile behaviour of micrometre-sized specimen fabricated from a metastable
austenitic stainless steel, Corros. Sci. 53 (2011) 529533.
[10] M. Koyama, E. Akiyama, T. Sawaguchi, K. Ogawa, I.V. Kireeva, Y.I. Chumlyakov,
K. Tsuzaki, Hydrogen-assisted quasi-cleavage fracture in a single crystalline
type 316 austenitic stainless steel, Corros. Sci. 75 (2013) 345353.
[11] R.J. Walter, W.T. Chandler: effects of high-pressure hydrogen on metals at
ambient temperature, Report no. NASA-CR-102425, Rocketdyne, Canoga Park,
CA, 1969.

H.F. Jackson et al. / Corrosion Science 77 (2013) 210221


[12] H.G. Nelson, Hydrogen embrittlement, in: C.L. Briant, S.K. Banerji (Eds.),
Treatise on Materials Science and Technology: Embrittlement of Engineering
Alloys, Academic Press, New York, NY, 1983, pp. 275359.
[13] G.R. Caskey Jr., Hydrogen effects in stainless steels, in: R.A. Oriani, J.P. Hirth,
M. Smialowski (Eds.), Hydrogen Degradation of Ferrous Alloys, Noyes
Publications, Park Ridge, NJ, 1985, pp. 822862.
[14] N.R. Moody, S.L. Robinson, W.M. Garrison Jr., Hydrogen effects on the
properties and fracture modes of iron-based alloys, Res. Mech. 30 (1990)
143206.
[15] C. San Marchi, B.P. Somerday: technical reference on hydrogen compatibility of
materials, Report no. SAND2008-1163, Sandia National Laboratories,
Livermore, CA, 2008. <http://www.sandia.gov/matlsTechRef/>.
[16] J.A. Brooks, A.W. Thompson, Microstructural development and solidication
cracking susceptibility of austenitic stainless steel welds, Int. Mater. Rev. 36
(1991) 1644.
[17] J.A. Brooks, A.J. West, Hydrogen induced ductility losses in austenitic stainless
steel welds, Metall. Trans. A 12A (1981) 213223.
[18] J.A. Brooks, A.J. West, A.W. Thompson, Effect of weld composition and
microstructure on hydrogen assisted fracture of austenitic stainless steels,
Metall. Trans. A 14A (1983) 7584.
[19] M.J. Morgan, G.K. Chapman, M.H. Tosten, S.L. West, Tritium effects on fracture
toughness of stainless steel weldments, in: Trends in welding research:
proceedings of the 7th international conference, ASM, International, 2005, pp.
743748.
[20] B.P. Somerday, M. Dadfarnia, D.K. Balch, K.A. Nibur, C.H. Cadden, P. Sofronis,
Hydrogen-assisted crack propagation in austenitic stainless steel fusion welds,
Metall. Mater. Trans. A 40A (2009) 23502362.
[21] H.F. Jackson, K.A. Nibur, C. San Marchi, J.D. Puskar, B.P. Somerday, Hydrogenassisted crack propagation in 304l/308l and 21cr6ni9mn/308l austenitic
stainless steel fusion welds, Corros. Sci. 60 (2012) 136144.
[22] K.A. Nibur, B.P. Somerday, D.K. Balch, C. San Marchi, The role of localized
deformation in hydrogen-assisted crack propagation in 21cr6ni9mn
stainless steel, Acta Mater. 57 (2009) 37953809.
[23] C. San Marchi, B.P. Somerday, J. Zelinski, X. Tang, G.H. Schiroky, Mechanical
properties of super duplex stainless steel 2507 after gas phase thermal
precharging with hydrogen, Metall. Mater. Trans. A 38A (2007) 27632775.
[24] W.C. Luu, P.W. Liu, J.K. Wu, Hydrogen transport and degradation of a
commercial duplex stainless steel, Corros. Sci. 44 (2002) 17831791.
[25] W.J. Mills, Fracture toughness of type 304 and 316 stainless steels and their
welds, Int. Mater. Rev. 42 (1997) 4582.
[26] F.P. Ford, The crack-tip system and its relevance to the prediction of cracking
in aqueous environments, in: R.P. Gangloff, M.B. Ives (Eds.), Environmentinduced Cracking of Metals, NACE, Houston, TX, 1990, pp. 139165.
[27] A.M. Brass, J. Chene, Hydrogen uptake in 316l stainless steel: Consequences on
the tensile properties, Corros. Sci. 48 (2006) 32223242.
[28] T. Zakroczymski, A. Glowacka, W. Swiatnicki, Effect of hydrogen concentration
on the embrittlement of a duplex stainless steel, Corros. Sci. 47 (2005) 1403
1414.
[29] J.R. Buckley, D. Hardie, The effect of pre-straining and d-ferrite on the
embrittlement of 304l stainless steel by hydrogen, Corros. Sci. 34 (1993) 93
107.
[30] C. San Marchi, T. Michler, K.A. Nibur, B.P. Somerday, On the physical
differences between tensile testing of type 304 and 316 austenitic stainless
steels with internal hydrogen and in external hydrogen, Int. J. Hydrogen
Energy 35 (2010) 97369745.
[31] P. Mllner, C. Solenthaler, P.J. Uggowitzer, M.O. Speidel, On the effect of
nitrogen on the dislocation structure of austenitic stainless steel, Mater. Sci.
Eng. A A164 (1993) 164169.
[32] L. Vitos, J.-O. Nilsson, B. Johansson, Alloying effects on the stacking fault energy
in austenitic stainless steels from rst-principles theory, Acta Mater. 54 (2006)
38213826.
[33] J. Talonen, H. Hnninen, Formation of shear bands and strain-induced
martensite during plastic deformation of metastable austenitic stainless
steels, Acta Mater. 55 (2007) 61086118.
[34] R.M. Latanision, J.A.W. Ruff, The temperature dependence of stacking fault
energy in FeCrNi alloys, Metall. Trans. 2 (1971) 505509.
[35] F. Lecroisey, A. Pineau, Martensitic transformations induced by plastic
deformation in the FeNiCrC system, Metall. Trans. 3 (1972) 387396.
[36] F. Abrassart, Stress-induced c ? a martensitic transformation in two carbon
stainless steels. Application to trip steels, Metall. Trans. 4 (1973) 22052216.

221

[37] J. Singh, Inuence of deformation on the transformation of austenitic stainless


steels, J. Mater. Sci. 20 (1985) 31573166.
[38] D.G. Ulmer, C.J. Altstetter, Hydrogen-induced strain localization and failure of
austenitic stainless steels at high hydrogen concentrations, Acta Metall. Mater.
39 (1991) 12371248.
[39] K.A. Nibur, D.F. Bahr, B.P. Somerday, Hydrogen effects on dislocation activity in
austenitic stainless steel, Acta Mater. 54 (2006) 26772684.
[40] ASTM, Standard test method for measurement of fracture toughness, ASTM
International, West Conshohocken, PA, 2009.
[41] ASTM, Standard test method for J-integral characterization of fracture
toughness, ASTM International, West Conshohocken, PA, 1996.
[42] C. San Marchi, B.P. Somerday, S.L. Robinson, Permeability, solubility and
diffusivity of hydrogen isotopes in stainless steels at high gas pressures, Int. J.
Hydrogen Energy 32 (2007) 100116.
[43] D.J. Kotecki, T.A. Siewert, Wrc-1992 constitution diagram for stainless steel
weld metals: a modication of the wrc-1988 diagram, Weld. J. 71 (1992) 171
179.
[44] C.D. Beachem, A new model for hydrogen-assisted cracking (hydrogen
embrittlement), Metall. Trans. 60 (1972) 437451.
[45] H.K. Birnbaum, P. Sofronis, Hydrogen-enhanced localized plasticity - a
mechanism for hydrogen-related fracture, Mater. Sci. Eng. A A176 (1994)
191202.
[46] D.P. Abraham, C.J. Altstetter, Hydrogen-enhanced localization of plasticity in
an austenitic stainless steel, Metall. Mater. Trans. A 26A (1995) 28592871.
[47] C.G. Rhodes, A.W. Thompson, The composition dependence of stacking fault
energy in austenitic stainless steels, Metall. Trans. A 8A (1977) 19011906.
[48] R.E. Schramm, R.P. Reed, Stacking fault energies of seven commercial
austenitic stainless steels, Metall. Trans. A 6A (1975) 13451351.
[49] P.L. Mangonon Jr., G. Thomas, The martensite phases in 304 stainless steel,
Metall. Trans. 1 (1970) 15771586.
[50] J.F. Breedis, L. Kaufman, Formation of hcp and bcc phases in austenitic iron
alloys, Metall. Trans. 2 (1971) 23592371.
[51] L. Rmy, A. Pineau, Observation of stacked layers of twins and e martensite in a
deformed austenitic stainless steel, Metall. Trans. 5 (1974) 963965.
[52] G. Olson, M. Cohen, A general mechanism of martensitic nucleation: Part i.
General concepts and the fcc ? hcp transformation, Metall. Trans. A 7A (1976)
18971904.
[53] L. Rmy, The interaction between slip and twinning systems and the inuence
of twinning on the mechanical behavior of fcc metals and alloys, Metall. Trans.
A 12A (1981) 387408.
[54] P. Mllner, C. Solenthaler, P.J. Uggowitzer, M.O. Speidel, Brittle fracture in
austenitic steel, Acta Metall. Mater. 42 (1994) 22112217.
[55] P. Mllner, On the ductile to brittle transition in austenitic steel, Mater. Sci.
Eng. A A234236 (1997) 9497.
[56] S. Takaki, T. Furuya, Y. Tokunaga, Effect of si and al additions on the low
temperature toughness and fracture mode of fe27mn alloys, ISIJ Int. 30
(1990) 632638.
[57] J. Brooks, J. Williams, A. Thompson, Microstructural origin of the skeletal
ferrite morphology of austenitic stainless steel welds, Metall. Trans. A 14A
(1983) 12711281.
[58] R. Ritchie, A. Thompson, On macroscopic and microscopic analyses for crack
initiation and crack growth toughness in ductile alloys, Metall. Trans. A 16A
(1985) 233248.
[59] A. Turnbull, Hydrogen diffusion and trapping in metals, in: R.P. Gangloff, B.P.
Somerday (Eds.), Gaseous Hydrogen Embrittlement of Materials in Energy
Technologies, Woodhead Publishing, Oxford, UK, 2012, pp. 89128.
[60] M.I. Luppo, A. Hazarabedian, J. Ovejero-Garcia, Effects of delta ferrite on
hydrogen embrittlement of austenitic stainless steel welds, Corros. Sci. 41
(1999) 87103.
[61] J.P. Hirth, Effects of hydrogen on the properties of iron and steel, Metall. Trans.
A 11A (1980) 861890.
[62] T.-P. Perng, C.J. Altstetter, Effects of deformation on hydrogen permeation in
austenitic stainless steels, Acta Metall. 34 (1986) 17711781.
[63] M.R. Louthan Jr., R.G. Derrick, Hydrogen transport in austenitic stainless steel,
Corros. Sci. 15 (1975) 565577.
[64] M. Nagumo, Hydrogen related failure of steels a new aspect, Mater. Sci.
Technol. 20 (2004) 940950.
[65] T. Neeraj, R. Srinivasan, J. Li, Hydrogen embrittlement of ferritic steels:
observations on deformation microstructure, nanoscale dimples and failure by
nanovoiding, Acta Mater. 60 (2012) 51605171.

Potrebbero piacerti anche