2nd Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

International Journal of Fatigue 104 (2017) 158–170

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Fatigue crack non-propagation assisted by nitrogen-enhanced


dislocation planarity in austenitic stainless steels
Kishan Habib, Motomichi Koyama ⇑, Toshihiro Tsuchiyama, Hiroshi Noguchi
Kyushu University, 744 Motooka, Nishi-ku, Fukuoka 819-0395, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Rotating bending fatigue tests were conducted to assess the fatigue crack propagation behavior of the Fe-
Received 23 May 2017 25Cr-1N and Fe-18Cr-14Ni austenitic steels in terms of the microstructure, crack propagation paths, and
Received in revised form 15 July 2017 non-propagating fatigue crack characteristics. The Fe-25Cr-1N steel exhibited a non-propagating fatigue
Accepted 17 July 2017
crack at the fatigue limit (310 MPa), but this did not occur in the Fe-18Cr-14Ni steel at the fatigue limit
Available online 19 July 2017
(110 MPa). The non-propagating fatigue crack observed in the Fe-25Cr-1N steel was produced by
roughness-induced crack closure. This phenomenon was caused by the enhanced planar dislocation
Keywords:
and high dislocation pile-up stress resulting from the suppression of cross-slip, which inhibited the dis-
High nitrogen austenitic steel
Fatigue crack growth
location emission from the crack tip. The Fe-25Cr-1N steel exhibited a lower fatigue crack growth rate
Dislocation planarity than the Fe-18Cr-14Ni steel because of the enhanced dislocation planarity produced by the Cr-N interac-
High cycle fatigue tion. The Cr-N interaction affected the fatigue crack growth behavior as follows. The short crack region
exhibited a planar glide dislocation pattern, but multiple slip systems were activated as the crack length-
ened. As the dislocation pattern remained planar on each slip plane, the crack propagation occurred along
the {1 1 1}c slip planes, even in the long crack. Moreover, the dislocation pile-up at the grain boundaries
caused grain boundary subcracks, which can induce crack toughening through mechanisms such as stress
redistribution. These positive effects contributed to the lower fatigue crack growth rate in the Fe-25Cr-1N
steel than the Fe-18Cr-14Ni steel.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction planarity, resulting in an increase in the work hardening capacity


and improved strength and elongation characteristics [11].
Austenitic stainless steels are widely used in industrial applica- In addition to the mechanical properties under monotonic load-
tions including cryogenic technology, nuclear equipment, biomed- ing, the fatigue resistance is the most crucial mechanical property
ical devices, and structural components owing to their excellent for structural design. Fatigue-related phenomena are a major cause
mechanical properties and corrosion resistance [1–4]. In terms of of accidental failures in structural applications, making it impor-
the alloy design of advanced austenitic steels, the addition of solute tant to know the fatigue limit and understand its underlying mech-
nitrogen plays a key role in austenite stabilization, improving the anisms to ensure safe structural design incorporating high nitrogen
mechanical properties and enhancing the corrosion resistance of steels for long-life structures. The fatigue properties of nitrogen-
austenitic stainless steels [5–9]. When the nitrogen content of an alloyed austenitic steel have previously been investigated, with
alloy exceeds 0.4 wt%, it is referred to as a high nitrogen austenitic reports of a superior fatigue resistance [12,13]. From the fatigue
stainless steel. Interstitial nitrogen introduces large elastic distor- limit point of view, the crack initiation and non-propagating crack
tions to the crystal lattice, resulting in solid solution hardening. phenomena are major factors in determining the fatigue strength
Furthermore, it has been argued that the multiple positive effects [14]. Factors controlling the crack resistance in fatigue crack non-
of nitrogen result from an attractive atomic interaction between propagation include strain aging at a mechanically small crack
Cr and N and not simply as an effect of nitrogen alone. Specifically, tip [15–17], crack closure mechanisms such as plasticity induced
dislocation slip causes decomposition of the Cr-N couples, which crack closure [18], and branching at the crack tip [19]. In addition,
induces local softening [10]. This effect enhances the dislocation the crack propagation path and subcrack interactions also affect
the fatigue crack growth resistance [20,21]. However, to the best
of our knowledge, a comprehensive understanding of the contribu-
⇑ Corresponding author. tion of these factors in high nitrogen steels has not yet been ade-
E-mail address: koyama@mech.kyushu-u.ac.jp (M. Koyama). quately achieved.

http://dx.doi.org/10.1016/j.ijfatigue.2017.07.019
0142-1123/Ó 2017 Elsevier Ltd. All rights reserved.
K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170 159

The nature of fatigue crack resistance in metals is defined by 14Ni steels were reported as 22 [25] and 35 mJ/m2 [26],
dislocation motion at the crack tip and the associated plastic zone respectively.
evolution. The characteristic dislocation motion of the high nitro- The fatigue test specimens shown in Fig. 1 were cut by EDM,
gen steels, namely the planar dislocation glide, must therefore and the damage layer was subsequently removed by mechanical
affect the crack growth resistance and fatigue limit. At a minimum and electrochemical polishing. A hole with both diameter and
we would expect that a strong dislocation pile-up stress arising depth equal to 100 lm was drilled into the center of the specimen
from the enhanced dislocation planarity would prevent dislocation on the top surface, in order to serve as a controlled crack initiation
emission from the crack tip, thus enhancing the fatigue crack site. The fatigue test specimen was fixed on a round bar with the jig
growth resistance. Furthermore, a heterogeneous dislocation array configured as reported in our previous paper [26]. The fatigue tests
can alter the evolution behavior in the plastic zone. were performed on an Ono-type rotating bending fatigue test
A further possible effect of the nitrogen-related dislocation pla- machine at room temperature, with a stress ratio of 1 and fre-
narity occurs in the crack propagation path. The dislocation planar quency 30 Hz. The fatigue cracks were observed using an optical
array, which is maintained even at a high stress, can change the microscope using the replica technique, and the fracture surface
crack propagation path [22] and cause crack branching and deflec- by scanning electron microscopy (SEM) at an accelerating voltage
tion. For example, the dislocation planar array has been reported to of 15 kV. The fractured specimens were mechanically polished
cause cracking along a slip plane [23]. Furthermore, the high pile- along an edge and electron backscatter diffraction (EBSD) measure-
up stress arising from the planarity may cause intergranular crack ments were performed at a distance of 100 lm (short crack) and
initiation and propagation. Therefore, considering at a minimum 400 lm (long crack) from the bottom of the hole, with a beam step
the required crack opening stress and propagation path, the size of 50 nm and an accelerating voltage of 20 kV. Electron chan-
nitrogen-enhanced dislocation planarity is strongly expected to neling contrast (ECC) imaging was carried out at an accelerating
positively alter the fatigue crack growth resistance. This study voltage of 30 kV. These SEM-based analyses were conducted for
therefore assesses the fatigue crack propagation characteristics of the sample cross-section to observe microstructures underneath
a high nitrogen Fe-Cr steel under the influence of enhanced dislo- a fatigue crack that propagated in a plain strain condition. Crack
cation planarity. growth in a plane strain condition is the primary phenomenon in
fatigue, although crack growth behavior in a plane stress condition,
i.e. surface crack growth, has a correlation with that in the corre-
2. Experimental procedure sponding plain strain condition.

Two steel compositions were used in this study: Fe-25Cr-1N


and Fe-18Cr-14Ni (mass%). The detailed chemical composition of 3. Experimental results
each steel is listed in Table 1. The Fe-25Cr-1N steel specimen
was prepared by solution nitriding with the base material of Fe- 3.1. Initial microstructure and tensile properties
25Cr (mass%) alloy a single Fe phase. The ingot was produced by
induction melting in a vacuum, homogenization at 1473 K, and Fig. 2(a) and (b) confirm that the Fe-18Cr-14Ni and Fe-25Cr-1N
hot-rolling to produce a 10 mm thick plate. The fatigue test speci- steels had a fully austenitic microstructure before fatigue testing.
men was cut from the ingot and solution-nitriding was performed The average grain sizes observed in the Fe-18Cr-14Ni and Fe-
for the fatigue specimen at 1473 K for 86.4 ks in a nitrogen gas 25Cr-1N steels were measured to be 42 and 22 lm respectively,
atmosphere at 0.1 MPa, followed by water quenching to ensure including annealing twin boundaries.
that no austenite decomposition or nitride precipitation occurred Fig. 3 shows the engineering stress-strain (S-S) curves of the Fe-
on cooling. As coarse-grained high nitrogen steels can exhibit brit- 18Cr-14Ni and Fe-25Cr-1N steels. The Fe-18Cr-14Ni steel exhibited
tle intergranular cracking, grain refinement was carried out on the a lower work hardening capacity and tensile strength than those of
solution-nitrided specimen by heat treatment in two stages [24]. the Fe-25Cr-1N steel. The 0.2% proof stress, tensile strength, and
The specimen was first subjected to isothermal heat treatment at elongation are listed in Table 2.
1173 K for 0.42 ks, followed by re-austenitization at 1473 K for
0.3 ks and water cooling. The specimen was then mechanically
and electrochemically polished to remove the damage layer before 3.2. Fatigue life and relevant crack growth characteristics
fatigue testing. The tensile properties of the Fe-25Cr-1N steel have
been reported to be improved as a result of grain refinement [24]. Fig. 4(a) shows a plot of the number of cycles to failure against
The Fe-18Cr-14Ni ingot was prepared by vacuum induction stress amplitude for the Fe-25Cr-1N and Fe-18Cr-14Ni steels. The
melting, forging and groove-rolling at 1273 K. The hot-rolled bar stress amplitude, which does not result in failure at 2  107 cycles,
was then solution treated at 1273 K for 1 h, followed by water is referred as the fatigue strength in this study. The fatigue
quenching to suppress the uncontrolled formation of second strengths of the Fe-25Cr-1N and Fe-18Cr-14Ni steels were
phases and segregation. The fatigue test specimen was cut by elec- 310 MPa and 110 MPa, respectively. As differences in the fatigue
tric discharge machining (EDM), and the damage layer was limit and fatigue life are generally explained by the hardness or
removed by mechanical and electrochemical polishing. tensile strength, a plot of the stress amplitude (normalized by
The tensile tests for both steels were conducted on three spec- the tensile strength) against the number of cycles to failure is
imens with a gauge length of 30 mm, 4 mm width, and 1 mm shown in Fig. 4(b). The fatigue strength of the Fe-25Cr-1N steel is
thickness, at room temperature and an initial strain rate of 10 3/ clearly higher than that of the Fe-18Cr-14Ni steel, even when nor-
s. The stacking fault energies of the Fe-25Cr-1N and Fe-18Cr- malized by the tensile strength. In contrast, the fatigue life of the

Table 1
Chemical compositions of the materials used in this study (mass%).

Steel Ni Mn Cr Si C P S N O Al Fe
Fe-25Cr-1N – <0.01 25.4 0.03 0.002 0.006 0.0005 1.0 0.0036 0.039 Bal
Fe-18Cr-14Ni 14.6 <0.001 18.8 <0.001 0.002 <0.001 0.0014 – – – Bal
160 K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170

Fig. 1. Schematic showing a fatigue test specimen with dimensions (in mm).

Fig. 2. Optical micrographs showing the microstructures of the (a) Fe-18Cr-14Ni and (b) Fe-25Cr-1N steels.

Fig. 3. Engineering stress-strain curves for the Fe-18Cr-14Ni and Fe-25Cr-1N steels.

Table 2
The 0.2% proof stress, tensile strength, and elongation of the Fe-25Cr-1N and
Fe-18Cr-14Ni steels.

Steel 0.2% Proof Tensile strength Elongation


stress (MPa) (MPa) (%)
Fe-25Cr-1N 950 ± 14 1200 ± 9 41 ± 1
Fe-18Cr-14Ni 290 ± 7 520 ± 10 50 ± 1.5

Fe-25Cr-1N steel at higher normalized stress amplitudes is almost


identical to that of the Fe-18Cr-14Ni steel. Fig. 4. (a) Stress amplitude versus fatigue life for the Fe-25Cr-1N and Fe-18Cr-14Ni
Fig. 5 shows the crack growth curves of the two steels at the steels. (b) Number of cycles to failure plotted against stress amplitude normalized
same stress amplitude to tensile strength ratio of 0.36. The fatigue by the tensile strength.
K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170 161

are aligned parallel to the main crack tip, as shown by red and
white arrows in Fig. 6(f1) and (g1).
Fig. 7 shows the SEM fracture surface images of the Fe-18Cr-
14Ni steel, where Fig. 7(a) gives an overview of the fracture sur-
face. Fig. 7(b) shows the crack initiation site from the hole, with
a flat surface around it. The early and intermediate propagation
regions exhibited surface features resembling a river pattern, as
shown in Fig. 7(c and d). The river pattern decreased as the crack
length increased (Fig. 7(e)), and ductile fatigue striations appeared
with increasing crack length as indicated by red arrows in Fig. 7(f).
The images in Fig. 8 labeled (a) show the results of the EBSD
measurements performed at a location 100 lm from the bottom
of the drill hole in the Fe-18Cr-14Ni fractured specimen. The EBSD
images were taken from the sample cross-section indicated by
arrows in Fig. 8(a1), and the corresponding cross-sectional SEM
image is shown in Fig. 8(a2). Fig. 8(a3) shows the inverse pole figure
map in the rolling direction occurring parallel to the loading direc-
tion (RD-IPF). The {1 1 1}c plane traces have been highlighted as
red lines in the RD-IPF map. Here, some parts of the crack surface
occur along the {1 1 1}c planes, as highlighted by yellow lines near
the fracture surface. Fig. 8(a4) shows a grain orientation deviation
angle (GROD) map, which indicates the degree of plastic deforma-
tion present. In the GROD map, the maximum deviation angle is
approximately 18° and the deviation angle distribution is hetero-
geneous. Fig. 8(a5) shows an image quality (IQ) map coupled with
an identification of the R3 twin boundaries. The IQ value deterio-
rates when two Kikuchi patterns associated with the parent and
twin crystals overlap. The reduction in IQ value depicts even thin
deformation twin plates [27–29]. The IQ map reveals only anneal-
ing twins, and no plate-like microstructure (such as deformation
twins) is observed.
The images in Fig. 8 labeled (b) show the results of the EBSD
measurements performed at a location 400 lm from the bottom
of the drill hole in the Fe-18Cr-14Ni fractured specimen. The EBSD
images were taken from the sample cross-section indicated by
arrows in Fig. 8(b1), and the corresponding cross-sectional SEM
image is shown in Fig. 8(b2). The {1 1 1}c plane traces are high-
lighted by red lines in the RD-IPF map (Fig. 8(b3)), indicating that
the crack surface in this region is not along the {1 1 1}c planes at
any point. The GROD map (Fig. 8(b4)) reveals that the maximum
Fig. 5. Fatigue crack growth rate curves for the two steels at the same stress ratio of
deviation angle is approximately 29°, which is markedly higher
0.36: (a) crack length versus number of cycles, (b) crack growth versus number of
cycles, and (c) crack growth versus crack length. than that of the short crack condition. In addition, the highly
deformed region is homogenous along the fracture surface. As in
Fig. 8(a5), no plate-like microstructure was observed in the corre-
sponding IQ map (Fig. 8(b5)).
crack growth for the Fe-25Cr-1N steel was slower compared to that Fig. 9 shows the electron channeling contrast (ECC) images of
of the Fe-18Cr-14Ni steel. A sudden increase in the crack growth the Fe-18Cr-14Ni steel. Fig. 9(a) reveals the dislocation structure
rate of the Fe-25Cr-1N steel due to the crack coalescence observed, at a crack length of 300 lm from the bottom of the drill hole,
indicated by an arrow in Fig. 5(b). The lower fatigue crack growth where the dislocations have become tangled with each other and
rate of the Fe-25Cr-1N steel results in a higher fatigue life at low a cellular dislocation type has formed. Fig. 9(b) shows the deforma-
stress amplitudes compared to the Fe-18Cr-14Ni steel (Fig. 4). tion twin plates (as indicated by yellow arrows) at a crack length of
600 lm from the bottom of the drill hole, which depend on the
crystallographic orientation. The twin plates are very thin (less
3.3. Microscopic features of crack growth and associated dislocation than 300 nm) compared to other steels.
patterns
3.3.2. Fe-25Cr-1N steel
3.3.1. Fe-18Cr-14Ni steel Fig. 10 shows the replica images of the Fe-25Cr-1N steel
Fig. 6 shows the replica images of the Fe-18Cr-14Ni steel obtained during fatigue testing at stress amplitude of 450 MPa. A
obtained during fatigue testing. A crack initiated at the left side fatigue crack was initiated at 5  104 cycles from the left side of
of the drilled hole at 1.5  105 cycles as shown in Fig. 6(c), while the drill hole (Fig. 10(b)) and 6  104 cycles from the right side
a crack initiated along the slip lines on the right side of the hole of the drill hole (Fig. 10(c)). The cracks subsequently propagated
at 5  104 cycles and propagated to meet it, as shown in Fig. 6 in a direction almost perpendicular to the loading direction on both
(d2–f2). The fatigue crack propagation was subsequently near- sides of the drill hole as shown in Fig. 10(d–g). Many subcracks ini-
perpendicular to the loading direction as shown in Fig. 6(b–g). tiated near the main crack indicated by red arrows in Fig. 10(d1,d2,
Intensive slip bands were formed around the crack (Fig. 6(d1– e2,f1,g1), and some coalesced with it indicated by yellow arrows in
g1)), resulting in crack branching. The tips of the crack branches Fig. 10(e1) and (f2). Crack branching occurred at one point, as
162 K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170

Fig. 6. Optical micrographs of the Fe-18Cr-14Ni steel during fatigue testing at a stress amplitude of 150 MPa: (a) 0, (b) 5  104, (c) 1.5  105, (d) 2.9  105, (e) 3.5  105, (f)
3.8  105, and (g) 4.8  105 cycles. The red and white arrows indicate the main crack and subcracks, respectively. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)
K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170 163

Fig. 7. Scanning electron microscopy (SEM) images of the fatigue fracture surface for the Fe-18Cr-14Ni steel tested at 150 MPa: (a) overview of the fracture surface, (b) crack
initiation site from the drill hole, (c) early and (d, e) intermediate propagation regions, and (f) late propagation regions. The locations of images (b-e) are indicated in part (a),
outlined by dashed lines. The small red arrows in part (f) indicate ductile fatigue striations. The inset of (f) indicates a magnified image of the region outlined by yellow line.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

shown by white and blue arrows in Fig. 10(g2). The crack propaga- images were taken from the sample cross-section indicated by
tion was rigorous, exhibiting a zigzag pattern as observed in Fig. 10 arrows in Fig. 13(b1), and the corresponding cross-sectional SEM
(e1) and (g2). Compared with the Fe-18Cr-14Ni steel, subcracks ini- image is shown in Fig. 13(b2). The RD-IPF map (Fig. 13(b3)) shows
tiated at many points in the Fe-25Cr-1N steel and the subsequent the {1 1 1}c plane traces highlighted by red lines; they indicate that
crack coalescence and branching were the main attributes of its the crack surface in this region also appeared partially along the
surface crack propagation behavior. {1 1 1}c planes. The GROD map (Fig. 13(b4)) gives a maximum
Fig. 11 shows the replica images of a non-propagating fatigue deviation angle of approximately 9°, which is almost identical to
crack in the Fe-25Cr-1N steel at a stress amplitude of 310 MPa. the value obtained from the short crack. No plate-like microstruc-
The crack initiated at 4.5  106 cycles from the right side of the ture was in the corresponding IQ map, as shown in Fig. 13(b5).
drill hole as shown in Fig. 11(b). The crack did not propagate any Fig. 14 shows the ECC images of the Fe-25Cr-1N steel. Fig. 14(a)
further until 2  107 cycles (Fig. 11(e)), and no other cracks initi- shows the dislocation structure at the short crack (100 lm from
ated (Fig. 11(a–e)). the bottom of the drill hole), which is controlled by planar glide
Fig. 12 shows the SEM fractographs obtained from the Fe-25Cr- as shown in Fig. 14(a1). As the crack lengthens (500 lm from the
1N steel, where Fig. 12(a) gives an overview of the fracture surface. bottom of the drill hole, Fig. 14(b)), multiple slip systems are acti-
Fig. 12(b) indicates the crack initiation site, with striation-like fea- vated as shown in Fig. 14(b1). However, it should be noted that the
tures observed in the vicinity of the hole (shown by red arrows) pattern of dislocations on each slip plane remains planar, even in
along with flat surface features (indicated by white arrows). the long crack region.
Fig. 12(c) shows the early propagation region, which contain sub- Fig. 15 shows a further ECC image of the Fe-25Cr-1N steel at a
cracks and striations as indicated by red arrows in the magnified grain boundary located far from the fracture edge. The image
image (Fig. 12(c1)). As the crack length increased, the intermediate shows the pile-up of planar dislocations at the grain boundary.
propagation region revealed that the fracture surface was inter- The local contrast variation near the grain boundary (outlined by
granular (Fig. 12(d)), which increased as shown in the late propa- dashed yellow lines) indicates a large stress concentration at the
gation region (Fig. 12(e)). boundary, due to the elastic field associated with the dislocation
The images in Fig. 13 labeled (a) show the results of the EBSD pile-up.
measurements performed at a location 100 lm from the bottom
of the drill hole in the Fe-25Cr-1N fractured specimen. The EBSD
images were taken from the sample cross-section indicated by 4. Discussion
arrows in Fig. 13(a1), and the associated cross-sectional SEM image
is shown in Fig. 13(a2). In the RD-IPF map (Fig. 13(a3)), the {1 1 1}c The S-S curve shown in Fig. 3 reveals that the tensile strength of
plane traces have been highlighted by red lines; the map reveals the Fe-25Cr-1N steel is superior to that of the Fe-18Cr-14Ni steel.
that in some regions the crack surface did occur along the The increase in the tensile strength of the Fe-25Cr-1N steel is
{1 1 1}c planes, indicated by yellow lines near the fracture surface. mainly attributed to the solid solution and grain boundary harden-
The GROD map (Fig. 13(a4)) gives a maximum deviation angle of ing that occur when the N is in the interstitial solid solution [5,6]
approximately 10°, and shows that the deviation angle distribution and also to the grain refinement of the Fe-25Cr-1N steel [24].
is heterogeneous. Plate-like microstructure was not observed in Accordingly, the fatigue strength of the Fe-25Cr-1N steel was
the corresponding IQ map, as shown in Fig. 13(a5). higher than that of the Fe-18Cr-14Ni steel (Fig. 4(a)). However,
The images in Fig. 13 labeled (b) show the results of the EBSD the difference in the fatigue strength between the two steels
measurements performed at a location 400 lm from the bottom cannot be explained by the tensile strength variation alone
of the drill hole in the Fe-25Cr-1N fractured specimen. The EBSD (Fig. 4(b)). The difference in the fatigue resistance of the
164 K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170

Fig. 8. Cross-sectional electron backscatter diffraction (EBSD) analysis of the Fe-18Cr-14Ni steel. Analysis at the short crack (100 lm from the bottom of the drill hole) is
indicated by the label (a), and at the long crack (400 lm from the bottom of the drill hole) by the label (b): (a1 and b1) fracture surfaces, (a2 and b2) scanning electron
microscopy (SEM) images, (a3 and b3) image quality maps in the rolling direction occurring parallel to the loading direction(RD IPF), (a4 and b4) grain reference orientation
deviation angle (GROD) maps, and (a5 and b5) image quality maps with an identification of the R3 twin boundaries. The yellow lines in (a3) indicate the regions where the
crack surface is along one of the {1 1 1}c planes. The red lines in (a5) and (b5) indicate the R3 twin boundaries. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

Fe-25Cr-1N and Fe-18Cr-14Ni steels can be interpreted using the the Fe-18Cr-14Ni steel did not. In order to understand the charac-
characteristics of the fatigue crack growth behavior shown in teristics of the non-propagating fatigue crack behavior and
Fig. 5. Furthermore, the Fe-25Cr-1N steel exhibited non- enhanced crack growth resistance of the Fe-25Cr-1N steel, we dis-
propagating fatigue crack behavior at the fatigue limit, whereas cuss the fatigue crack behavior in detail in the following sections.
K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170 165

Fig. 9. Electron channeling contrast (ECC) images of the Fe-18Cr-14Ni steel: (a) dislocation cell structure at a crack length of 300 lm from the bottom of the hole, and (b)
deformation twin plates at a crack length of 600 lm from the bottom of the hole. The yellow lines indicate twin plates. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

4.1. Fatigue-induced dislocation patterns causing crack initiation (b)). Accordingly, typical fatigue striation features were observed
in Fig. 7(f) as a result of the large stress concentration and the acti-
Crack initiation occurred in the Fe-18Cr-14Ni steel along the vation of multiple slip systems and cross-slip with increasing crack
slip bands, as shown in Fig. 6(d2). The damage process began in length.
regions of localized cyclic strain, resulting in the formation of per- In contrast, the Fe-25Cr-1N steel exhibited subcrack formation,
sistent slip bands (PSBs). This process involves large extrusions and as shown in Fig. 10. The number of subcracks was significantly
intrusions with a high stress intensity factor, which cause cracks to higher than was observed in the Fe-18Cr-14Ni steel. Subcrack for-
initiate from the PSBs and encourage subsequent propagation mation can be caused by the enhancement of dislocation planarity
[30,31]. Generally, the dislocation arrangement in the matrix and by the Cr-N interaction (as mentioned in the previous section).
PSBs is strongly affected by the character of the slip. The slip char- Specifically, the Cr-N interaction suppresses cross-slip and
acter typically depends on the stacking fault energy, and the dislo- enhances planar slip, leading to subcrack development on the sur-
cation planarity is enhanced with decreasing stacking fault energy, face [37]. Subcrack formation has both positive and negative
which tends to initiates fatigue cracks at the slip bands [32]. This effects on the growth of fatigue cracks. Subcrack coalescence with
phenomenon would cause slip plane crack growth in the Fe- the main crack can accelerate fatigue crack growth, as shown in
18Cr-14Ni steel, which has a stacking fault energy of 35 mJ/m2. Figs. 5(b) and 10(e1) and (f2). However, a stress distribution near
However, we can also note that the crack along the slip plane the crack tip [20] and roughness-induced crack closure [38,39]
was deflected (even in the first grain from the drill hole) as shown can decelerate the crack growth. Specifically, stress release occurs
in Fig. 6(d1). This behavior can be caused by the formation of dis- when a subcrack is aligned parallel to the main crack [20], decreas-
location substructures such as cells, as observed in Fig. 9(a). ing the driving force for crack growth. The roughness-induced
In the Fe-25Cr-1N steel, fatigue cracks were frequently initiated crack closure phenomenon results from zigzag crack propagation,
along the grain boundaries (Fig. 10(d1) and (d2)). According to the through crack coalescence, and deflection as shown in Fig. 10(e1,
dislocation image shown in Fig. 15, the fatigue-induced disloca- f2,g1,g2).
tions exhibit planar features, which causes a dislocation-pile up
stress at the grain boundary. This grain boundary stress concentra- 4.3. Effects of crack length on the crystallographic features of crack
tion is a cause of the intergranular crack initiation observed in growth and subcrack formation
Fig. 12(d) and (e). Transgranular crack initiation along the slip
plane was also observed in some areas (Fig. 10(f1) and (g1)), and The microstructural fatigue crack propagation in the Fe-18Cr-
is associated with the dislocation planarity arising from the effect 14Ni steel occurred along the {1 1 1}c planes at some locations in
of the Cr-N couples. However, transgranular crack initiation the short crack region, as indicated by yellow lines in Fig. 8(a3).
occurred less frequently than intergranular crack initiation. However, the fatigue crack propagation features along the
{1 1 1}c planes disappeared in the long crack region (Fig. 8(b3)).
4.2. Micro-mechanisms of crack growth: crack branching and As no deformation twins were observed near the fracture surface
coalescence in the short crack region (Fig. 8(a5)), the crack propagation along
the {1 1 1}c planes cannot be attributed to deformation twinning.
The most important feature of crack propagation in the Fe- Propagation along the {1 1 1}c planes typically occurs as a result
18Cr-14Ni steel was the formation of intensive slip bands, crack of the accumulation of dislocations on a single {1 1 1}c plane
branching, and subcrack formation, as shown in Fig. 6. The [32,40], which contributes to the formation of damage sources
observed fatigue crack propagated continuously within the grain such as vacancies and dislocation dipoles during the fatigue pro-
interior, but crack branching occurred when it met the grain cess. However, as the maximum GROD value increased from the
boundary or deformation-induced bands and cells, as shown in short crack to the long crack (Fig. 8(a4) and (b4)), the plastic zone
Fig. 6(f1). Crack branching also occurred along internal microstruc- size and maximum plastic strain at the crack tip increased with
ture bands such as deformation twins, as shown in Fig. 6(g1). Pre- increasing crack length. This predominantly acts to activate
vious research has described the effect of microstructural features another {1 1 1}c plane, because of work hardening in the primary
including PSBs [33], cells [34], and twins [35] on crack propagation slip plane. The crystallographic features of the fatigue crack prop-
in austenitic steels. Hence, the occurrence of crack branching along agation path in the long crack region have therefore been blurred
the grain boundaries and deformation twins is a common crack out, as shown in Fig. 8(b3). Additionally, in the long crack region
propagation behavior in austenitic steels with a low stacking fault the large plastic strain at the crack tip created a stretching zone,
energy [36]. When the crack lengthened, the microstructure/crys resulting in the formation of ductile fatigue striations [41] as
tallography-dependent propagation behavior disappeared (Fig. 8 observed in Fig. 7(f).
166 K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170

Fig. 10. Optical micrographs of the Fe-25Cr-1N steel replicas taken during fatigue testing at 450 MPa: (a) 0, (b) 5  104, (c) 6  104, (d) 7  104, (e) 1  105, (f) 1.15  105, and
(g) 1.2  105cycles. The red arrows indicate subcracks and the yellow arrow indicates a point of crack coalescence. The blue and white arrows in image (g2) indicate the main
crack and a branched crack tip, respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170 167

Fig. 11. Optical micrographs of a non-propagating crack in the Fe-25Cr-1N steel at 310 MPa: (a) 0, (b) 4.5  106, (c) 5.3  106, (d) 9.7  106, and (e) 2  107 cycles. The
corresponding magnified images are shown in (c1), (d1), and (e1).

Fig. 12. Scanning electron microscopy (SEM) images of the fatigue fracture surface for the Fe-25Cr-1N steel tested at 450 MPa: (a) overview of the fracture surface, (b) site of
the crack initiation emanating from the drill hole, (c) early, (d) intermediate, and (e) late propagation regions as indicated by dashed lines in (a). Striation features are
indicated by red arrows, and flat regions by white arrows. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)

In contrast, the Fe-25Cr-1N steel exhibited fatigue crack propa- long crack (Fig. 13(b1)). This result is completely different to the
gation along the {1 1 1}c planes in both the short and long crack behavior observed in the Fe-18Cr-14Ni steel.
regions, as shown in Fig. 13(a3) and (b3). The occurrence of crack As seen in Fig. 13(a4), the degree of strain evolution/localization
propagation along the {1 1 1}c planes in both the long and short in the Fe-25Cr-1N steel was small compared to the Fe-18Cr-14Ni
crack regions is the major difference between the fatigue crack steel. This can be explained by two factors. The first relates to
propagation behavior of the Fe-18Cr-14Ni and Fe-25Cr-1N steels. the subcrack formation and subsequent crack coalescence. When
This behavior can mainly be attributed to the enhanced dislocation the crack growth occurs via crack coalescence, the plastic strain
planarity associated with the Cr-N interaction, which leads to sub- around the crack surface corresponds to a critical strain associated
crack formation. Fig. 14 shows the dislocation planarity at both the with the crack initiation, which must be lower than the plastic
short and long cracks; at the short crack (Fig. 14(a)) the disloca- strain evolution arising from the crack opening for growth result-
tions have a simple planar array, but at the long crack (Fig. 14 ing in a significant crack length to occur. Second, the enhanced dis-
(b)) multiple glide systems are activated. Note that the slip charac- location planar array in the vicinity of the crack surface facilitates
ter remains planar on each slip plane, resulting in the formation of mode II fatigue crack growth; this refers to the crack propagation
subcracks (even at the long crack) as shown in Fig. 10(g1) and (g2). path occurring along the {1 1 1}c planes, even in the long crack
This is due to the dislocation pile-up at the grain boundary, and as region. The dislocation planar glide-driven crack growth causes
a result the fracture surface exhibits intergranular fracture at the highly localized dislocation accumulation along the slip plane.
168 K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170

Fig. 13. Electron backscatter diffraction (EBSD) images of the Fe-25Cr-1N steel. Analysis at the short crack (100 lm from the bottom of drill hole) is indicated by the label (a),
and at the long crack (400 lm from the bottom of drill hole) by the label (b): (a1 and b1) fracture surface, (a2 and b2) scanning electron microscopy (SEM) image, (a3 and b3)
image quality maps in the rolling direction occurring parallel to the loading direction (IQ + RD IPF), (a4 and b4) grain reference orientation deviation angle (GROD) maps, and
(a5 and b5) IQ map with identification of the R3 twin boundaries. The yellow lines in (a3) indicate regions where the crack surface follows one of the {1 1 1}c planes. The red
lines in (e) indicate the R3 twin boundaries. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Therefore, when a crack propagates along the {1 1 1}c slip plane, crack growth stemming from the dislocation planarity, which was
the plastic zone size around the crack surface must be small. Fur- maintained even at the long crack. From this viewpoint, the plastic
thermore, as shown in Fig. 13(b4), the plastic strain around the strain level would increase with crack length, for the most part.
crack surface did not increase with crack length. This can also be However, compared with the Fe-18Cr-14Ni steel (which exhibited
explained by the reasons mentioned above. In terms of crack coa- mode I crack growth in the long crack region), the dependence of
lescence, the strain level is predominantly controlled by the critical the crack length on the plastic zone size and detectable strain level
strain required for the crack initiation. We can consider the mode II must be lower. This is due to the residual dislocation planar array
K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170 169

Fig. 14. Electron channeling contrast (ECC) images of the Fe-25Cr-1N steel, showing dislocation structure at: (a) the short crack (100 lm from the bottom of the drill hole)
and (b) the long crack (500 lm from the bottom of the drill hole). Images (a1) and (b1) show magnifications of the short and long cracks.

interaction. This is hypothesized to play a major role in the fatigue


crack non-propagation behavior, in a similar manner to that dis-
cussed in relation to crack growth resistance in Section 4.2. Specif-
ically, the enhanced planar dislocation glide causes crack
branching, deflection, and subcrack formation, facilitating the
roughening of the crack surface. This effect enhances the
roughness-induced crack closure, which can improve the crack
non-propagation limit. In addition, the high dislocation pile-up
stress stemming from the suppression of cross-slip prevents dislo-
cation emission from the crack tip, which may also increase the
limit of crack non-propagation. As a result, a non-propagating fati-
gue crack was only observed in the Fe-25Cr-1N steel in this study.
Consequently, the non-propagation of the crack enhanced the fati-
gue limit and fatigue life at low stress amplitudes, as shown in
Fig. 4.

Fig. 15. An electron channeling contrast (ECC) image of a grain boundary in the Fe- 5. Conclusion
25Cr-1N fractured specimen. The outlined region shows the elastic contrast
associated with the dislocation pile-up stress concentration.
A comparative study of two steels, Fe-25Cr-1N and Fe-18Cr-
14Ni, was conducted to examine the effect of nitrogen-enhanced
observed in Fig. 14(b1), which allowed mode II crack growth to dislocation planarity on the fatigue resistance, specifically the fati-
occur even in the long crack region. gue crack growth behavior and non-propagating fatigue crack phe-
nomenon. From the results, the following conclusions can be
drawn:
4.4. Role of nitrogen-enhanced dislocation planarity in non-
propagating crack behavior (1) The fatigue limit and fatigue life of the Fe-25Cr-1N steel
were higher than those of the Fe-18Cr-14Ni steel at rela-
The experimental results revealed that the Fe-25Cr-1N steel tively low stress amplitudes, due to nitrogen-enhanced dis-
exhibited a non-propagating fatigue crack at the fatigue limit location planarity.
(Fig. 11), whereas the Fe-18Cr-14Ni steel did not. The austenite (2) The enhanced planar dislocation glide in the Fe-25Cr-1N
in both the Fe-25Cr-1N and Fe-18Cr-14Ni steels was stable, and steel caused crack branching, deflection, and subcrack for-
possessed characteristics typical of low stacking fault energy mate- mation along the grain boundaries or {1 1 1}c slip planes,
rials, such as twin and extended dislocations. However, a signifi- which facilitated the roughness-induced crack closure. As a
cant difference between the two steels was found in the result, a non-propagating fatigue crack appeared at the fati-
nitrogen-enhanced dislocation planarity. The dislocation planarity gue limit. A further reason for the fatigue crack non-
tends to depend on the stacking fault energy and the short-range propagation associated with the dislocation planarity was
ordering of solute atoms. However, the enhanced dislocation pla- the high dislocation pile-up stress stemming from the sup-
narity observed in the Fe-25Cr-1N steel is an exceptional result, pression of cross-slip, as it prevented dislocation emission
produced by the short-range ordering associated with the Cr-N from the crack tip.
170 K. Habib et al. / International Journal of Fatigue 104 (2017) 158–170

(3) In the long crack region, multiple dislocation glide systems [17] Ritchie R, Lankford J. Small fatigue cracks: a statement of the problem and
potential solutions. Mater Sci Eng 1986;84:11–6.
were activated, but the slip character remained planar on
[18] Dougherty J, Srivatsan T, Padovan J. Fatigue crack propagation and closure
each slip plane, resulting in crack branching, deflection, behavior of modified 1070 steel: experimental results. Eng Fract Mech
and subcrack formation (even at the long crack). 1997;56:167–87.
(4) The number of subcracks in the Fe-25Cr-1N steel was higher [19] Murakami Y, Takahashi K. Torsional fatigue of a medium carbon steel
containing an initial small surface crack introduced by tension-compression
than in the Fe-18Cr-14Ni alloy, which resulted from crack fatigue. Fatigue Fract Eng Mater Struct 1998;21:1473–84.
initiation caused by dislocation accumulation at the grain [20] Shum DK, Hutchinson JW. On toughening by microcracks. Mech Mater
boundary in the Fe-25Cr-1N steel. Subcrack formation has 1990;9:83–91.
[21] Ritchie R. Mechanisms of fatigue crack propagation in metals, ceramics and
both negative and positive effects on fatigue crack growth. composites: role of crack tip shielding. Mater Sci Eng, A 1988;103:15–28.
Coalescence with a main crack accelerates the fatigue crack [22] Blankenship C, Starke E. THE fatigue crack growth behavior of the Al—Cu—Li
growth, whereas roughness-induced crack closure and stress alloy weldalite 049. Fatigue Fract Eng Mater Struct 1991;14:103–14.
[23] Tanaka K, Mura T. A dislocation model for fatigue crack initiation. J Appl Mech
redistribution by crack/crack interactions decelerates the (Trans ASME), 48 (1981) 97–103.
fatigue crack growth. [24] Onomoto T, Terazawa Y, Tsuchiyama T, Takaki S. Effect of grain refinement on
tensile properties in Fe-25Cr-1N alloy. ISIJ Int 2009;49:1246–52.
[25] Ojima M, Adachi Y, Tomota Y, Katada Y, Kaneko Y, Kuroda K, et al. Weak beam
TEM study on stacking fault energy of high nitrogen steels. Steel Res Int
Acknowledgements 2009;80:477–81.
[26] Habib K, Koyama M, Noguchi H. Impact of Mn–C couples on fatigue crack
growth in austenitic steels: is the attractive atomic interaction negative or
This work was financially supported by JSPS KAKENHI (Grant
positive? Int J Fatigue 2017;Part 99(1):1–12.
number JP16H06365). [27] Barbier D, Gey N, Bozzolo N, Allain S, Humbert M. EBSD for analysing the
twinning microstructure in fine-grained TWIP steels and its influence on work
References hardening. J Microsc 2009;235:67–78.
[28] Jin J-E, Lee Y-K. Strain hardening behavior of a Fe–18Mn–0.6C–1.5 Al TWIP
steel. Mater Sci Eng, A 2009;527:157–61.
[1] Hasegawa M. Stainless steel handbook. Tokyo: Nikkan Kogyo Shimbun Ltd.; [29] Koyama M, Sawaguchi T, Tsuzaki K. TWIP effect and plastic instability
1973. condition in an Fe–Mn–C austenitic steel. ISIJ Int 2013;53:323–9.
[2] Davis JR. Stainless steels. ASM international; 1994. [30] Hunsche A, Neumann P. Quantitative measurement of persistent slip band
[3] Gardner L. The use of stainless steel in structures. Prog Struct Mat Eng profiles and crack initiation. Acta Metall 1986;34:207–17.
2005;7:45–55. [31] Lin MR, Fine ME, Mura T. Fatigue crack initiation on slip bands: theory and
[4] Gludovatz B, Hohenwarter A, Catoor D, Chang EH, George EP, Ritchie RO. A experiment. Acta Metall 1986;34:619–28.
fracture-resistant high-entropy alloy for cryogenic applications. Science [32] Hong SI, Laird C. Fatigue crack initiation and growth behavior of Cu-16 at.% A1
2014;345:1153–8. single crystals. Fatigue Fract Eng Mater Struct 1991;14:143–69.
[5] Simmons J. Overview: high-nitrogen alloying of stainless steels. Mater Sci Eng, [33] Zhang GP, Takashima K, Shimojo M, Higo Y. Fatigue behavior of microsized
A 1996;207:159–69. austenitic stainless steel specimens. Mater Lett 2003;57:1555–60.
[6] Reed RP. Nitrogen in austenitic stainless steels. JOM 1989;41:16–21. [34] Altenberger I, Scholtes B, Martin U, Oettel H. Cyclic deformation and near
[7] Berns H. High nitrogen steels. Manufacture and application of high nitrogen surface microstructures of shot peened or deep rolled austenitic stainless steel
steels. ISIJ Int 1996;36:909–14. AISI 304. Mater Sci Eng, A 1999;264:1–16.
[8] Baba H, Kodama T, Katada Y. Role of nitrogen on the corrosion behavior of [35] Niendorf T, Rubitschek F, Maier H, Niendorf J, Richard H, Frehn A. Fatigue crack
austenitic stainless steels. Corros Sci 2002;44:2393–407. growth—microstructure relationships in a high-manganese austenitic TWIP
[9] Karaman I, Sehitoglu H, Maier H, Chumlyakov Y. Competing mechanisms and steel. Mater Sci Eng, A 2010;527:2412–7.
modeling of deformation in austenitic stainless steel single crystals with and [36] Lukáš P, Kunz L. Cyclic slip localisation and fatigue crack initiation in fcc single
without nitrogen. Acta Mater 2001;49:3919–33. crystals. Mater Sci Eng, A 2001;314:75–80.
[10] Vogt JB, Messai A, Foct J. Sensitivity of a high nitrogen austenitic stainless steel [37] Karaman I, Sehitoglu H, Maier HJ, Chumlyakov YI. Competing mechanisms and
to fatigue crack initiation. ISIJ Int 1996;36:862–6. modeling of deformation in austenitic stainless steel single crystals with and
[11] Terazawa Y, Ando T, Tsuchiyama T, Takaki S. Relationship between work without nitrogen. Acta Mater 2001;49:3919–33.
hardening behaviour and deformation structure in Ni-free high nitrogen [38] Ritchie R, Suresh S. Some considerations on fatigue crack closure at near-
austenitic stainless steels. Steel Res Int 2009;80:473–6. threshold stress intensities due to fracture surface morphology. Metall Trans A
[12] Vogt J-B. Fatigue properties of high nitrogen steels. J Mater Process Technol 1982;13:937–40.
2001;117:364–9. [39] Gray G, Williams J, Thompson A. Roughness-induced crack closure: an
[13] Dai Q, Yuan Z, Chen X, Chen K. High-cycle fatigue behavior of high-nitrogen explanation for microstructurally sensitive fatigue crack growth. Metall
austenitic stainless steel. Mater Sci Eng, A 2009;517:257–60. Trans A 1983;14:421–33.
[14] Miller K. The short crack problem. Fatigue Fract Eng Mater Struct [40] Buchinger L, Cheng A, Stanzl S, Laird C. The cyclic stress—strain response and
1982;5:223–32. dislocation structures of Cu 16 at.% Al alloy III: single crystals fatigued at low
[15] Wilson D, Tromans J. Effects of strain ageing on fatigue damage in low-carbon strain amplitudes. Mater Sci Eng 1986;80:155–67.
steel. Acta Metall 1970;18:1197–208. [41] Laird C. The influence of metallurgical structure on the mechanisms of fatigue
[16] Oates G, Wilson D. The effects of dislocation locking and strain ageing on the crack propagation. In: Fatigue crack propagation, ASTM International; 1967.
fatigue limit of low-carbon steel. Acta Metall 1964;12:21–33.

You might also like