You are on page 1of 12

International Journal of Fatigue 124 (2019) 70–81

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Predicting crack initiation site in polycrystalline nickel through surface T


topography changes

Jalal Fathi Solaa, Randall Keltonb, Efstathios I. Meletisb, Haiying Huanga,
a
Department of Mechanical and Aerospace Engineering, University of Texas at Arlington, Box 19018, Arlington, TX 76019, USA
b
Department of Materials Science and Engineering, University of Texas at Arlington, Box 19031, Arlington, TX 76019, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Fatigue loading generates plastic deformation and strain localization on the surface of polycrystalline metals,
Crack initiation which are known to be the precursor for crack initiation. The plastic deformations manifest themselves in the
Polycrystalline form of surface roughening that can be characterized using surface profiling techniques. In this study, surface
Plastic deformation topology changes on the surface of a nickel specimen under cyclic loading are monitored and analyzed to in-
Surface topography
vestigate the correlation between surface topography changes and crack initiation. Defining the surface
Fatigue
roughness parameters as the damage index, an algorithm was formulated to predict the crack initiation location
well before the formation of a visible crack. In addition, the crack initiation time can also be determined based
on the saturation of the surface roughness damage index as well as the rebound of the surface height near the
crack initiation location.

1. Introduction on the sensitivity of the measurement technique as well as the fre-


quency of measurements that can be realistically carried out.
Fatigue is the main failure mode for many engineering structures. Despite these difficulties, past research works have identified that
Fatigue life typically is divided into three different stages: crack in- the formation of slip steps are the necessary condition for crack in-
cubation or nucleation where cyclic hardening or softening occurs, itiation. When metals are cyclically loaded and the critical resolved
crack initiation where small measurable crack forms, and crack pro- shear stress on a slip system is reached, dislocations move along the slip
pagation [1]. By the time a measurable crack emerges, up to 70% of the direction, leading to the formation of slip steps. During the reverse
fatigue life has been spent under Low Cycle Fatigue (LCF). For High loading, the slip does not recover fully and cyclic slip irreversibility
Cycle Fatigue (HCF), the spent fatigue life increases to 90% [2–4]. In occurs [8,9]. Subsequently, by applying more cyclic loading, the slip
addition, the mechanism that drives crack initiation, i.e. plastic de- continues on the same path resulting in surface roughening and strain
formation, also governs early crack propagation. A better under- localization which eventually leads to persistent slip bands (PSBs) for-
standing of crack initiation behavior, therefore, could lead to more mation. Based on these understandings, cyclic slip irreversibility and
reliable fatigue life prediction, early crack prevention, and better design critical accumulated slip were proposed as fatigue indicator parameters
of fatigue resistant super-alloys. Unfortunately, the crack initiation lo- (FIP) [1,10–12]. In these studies, a critical value of irreversibility was
cations are typically unknown prior to crack initiation. This makes in- defined as a necessary condition for crack initiation. However, this
vestigating crack initiation very tedious and time-consuming; multiple condition is not sufficient for crack initiation [13]. Alternatively, other
potential crack initiation sites have to be monitored closely over researchers attempted to predict the crack initiation time based on the
thousands of cycles. Besides, crack initiation in polycrystalline metals accumulated energy in the slip bands [14–16]. The proposed theoretical
depends strongly on the local microstructures, such as the grain or- models and simulations have provided physical insights into crack in-
ientation, the characteristic of neighboring grains, the existence of de- itiation phenomena. However, all of the proposed methods rely on
fects or particles, etc. [1,5–7]. As a result, predicting crack initiation crystal plasticity simulations that have many simplifications and as-
location remains an unsolved challenge, despite intensive studies on sumptions. For example, many simulations only considered a few spe-
crack initiation carried out over the past century. Determining the time cific slip systems or the surface grains because the crystallographic
at which the crack initiates is also problematic; crack initiation is de- features underneath the surface cannot be measured easily. As a result,
fined as the time when a measurable crack is detected, which depends the simulation results are only partly verified by experimental


Corresponding author at: 500 W. First Street, WH211, Arlington, TX 76019, USA. Tel.: +1 817 272 0563; fax: +1 817 272 5010.
E-mail address: huang@uta.edu (H. Huang).

https://doi.org/10.1016/j.ijfatigue.2019.02.027
Received 3 December 2018; Received in revised form 13 February 2019; Accepted 19 February 2019
Available online 27 February 2019
0142-1123/ © 2019 Elsevier Ltd. All rights reserved.
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

its simplicity and high temporal and spatial resolution. Unlike AFM,
Nomenclature
these optical techniques are non-contacting; they use interference of
light for measuring the surface topography in a wide height range. As
Sa Arithmetic average surface roughness
such, the field of view is larger and the effect of out-of-plane movement
St Maximum height of the profile
within the grain or across grain boundaries can be measured. Taking
advantage of this technique, it was discovered that the largest surface
displacements were concentrated on triple points and grain boundaries
observations, making them inaccurate in practice [17]. One challenge
[32] and the surface roughening was dominated by the out-of-plane
in validating these FIPs is that they involve out-of-plane strain com-
movement of the grains after the heights of the slip bands have satu-
ponents, which are difficult to measure using conventional non-de-
rated [33]. Muravsky [34] showed that the fatigue process zone (FPZ)
structive characterization tools. Even though Digital Image Correlation
can be calculated from surface topography measured using a two-step
(DIC) has been extensively studied for investigating strain hetero-
phase-shifting interferometry. In addition, researchers have also ana-
geneity around potential crack initiation sites [18–20], DIC, when
lyzed the surface topography images in an attempt to determine the
carried out in two-dimensions (2D), does not provide quantitative in-
crack initiation time [35–37]. These studies, however, performed the
formation about the out-of-plane strain components. These might be the
analysis over relatively large areas that contains hundreds of grains or
reason that no published work has successfully predicted the crack in-
more. Vladimirov et al. and Jha et al. [35,36] did not validate their
itiation location based on 2D DIC strain measurements.
findings using Scanning Electron Microscope (SEM) images while the
Surface profiling techniques provide a quantitative measurement of
SEM images in [37] contained multiple cracks that are about
the surface topography and thus are capable of measuring the out-of-
200–300 μm in length. None of these work tried to predict the crack
plane displacements, at least on the sample surface. The extrusion
initiation site before the crack emerges.
height, measured using an Atomic Force Microscopy (AFM), was used
In our previous study, we adopted the SWLI technique to investigate
as a parameter to quantitatively represent strain irreversibility [21–23].
surface topography changes around the tip of a micro-sensitive small
They found that crack initiated when the extrusion height, normalized
crack [38–40]. It was observed that when the crack was arrested near a
by the grain size, reached a threshold. This threshold was found to
triple point, substantial surface topography change occurred in the FPZ
varying with the material. However, other studies have shown that
and a narrow depression was formed in front of the crack, through
some grains did not have crack initiation even though the extrusion
which the crack eventually grew. Based on these observations, we in-
height has reached the threshold [13,24]. HO et al [25] evaluated
troduced the surface roughness parameters as damage indices and
different surface roughness parameters to characterize the surface to-
formulated an algorithm that enables us to predict the tortuous crack
pographies of eight grains with crack initiation and eight grains
path over several grains. We also discovered that there is a damage
without. They claimed that the maximum peak height Sp parameter is
index threshold for a given stress intensity factor (SIF), beyond which
the best for differentiating the grains with crack initiation from those
the crack releases from an arresting point. When the crack was arrested,
without. Unfortunately, these AFM studies only investigated slip bands
the area in front of the crack was observed to reduce in height before
inside grains. As such, intergranular crack initiations and the out-of-
the crack was released from arresting and rebounded after the release.
plane grain displacements were not considered. Due to the raster
Since crack initiation and small crack propagation share similar driving
scanning mode of AFMs, imaging a large area could be very time-con-
mechanism and behavior, these discoveries motivated us to investigate
suming. In addition, AFM is not capable of measuring the surface to-
the surface topography changes during crack initiation in the present
pography of the narrow intrusions [9]. To overcome this problem, some
study. The surface topography changes of a pure nickel sample under
researchers combined AFM with the replica technique so that the for-
cyclic loading were monitored at real-time using an SWLI and analyzed
mation of thin intrusions between extrusions can be characterized
to predict the crack initiation location as well as to determine the crack
[3,26–31]. Using the replica technique, the crack length and site can be
initiation time. We demonstrated that the surface roughness damage
successfully detected, however, this technique cannot be performed in-
index, formulated based on the surface topography changes, is capable
situ and is labor intensive. More recently the study of surface profile
of predicting the crack initiation location well before a visible crack was
changes using optical measurement techniques, such as scanning white
detected using SEM. In addition, the crack initiation time was de-
light interferometry (SWLI), scanning laser confocal microscope (SLCM)
termined based on the saturation of the surface roughness damage
and two-step phase shifting interferometry has become popular due to
index as well as the recovery of the surface height near the crack

Fig. 1. (a) Dimensions of test specimen (all dimensions are in mm) (b) Test setup involving Electroforce fatigue testing machine, integrated with a scanning
whitelight interferometer (SWLI) surface profilometer.

71
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

when the sample was unloaded, i.e. the applied load was at zero, be-
cause of two reasons: (1) the entire system was slightly vibrating under
loading, which made acquiring clear SWLI images impossible; (2) the
imaging process could take up to 20 min and the sample is unloaded to
eliminate any possibility of relaxation. The sample was illuminated
using the built-in whitelight source of the SWLI. No additional lighting
was needed. The SWLI profilometer has two lenses of different magni-
fications; the 10× lens has a field of view (FOV) of 627 × 470 µm2 and
the 50× lens has an FOV of 235 × 175 µm2. The accompanying soft-
ware, vision64, enables acquiring multiple images by transversing the
lens in a raster motion and stitching them to form an image that is much
larger than the FOV of the imaging lens. More detailed information
about the operation of the SWLI can be found in the manufacturer’s
website [41].
Fig. 2. The region of interest defined after applying the first fatigue cycle. The The cyclic loading with a maximum load of 2900 N and a load ratio
surface topography image was acquired using a 10× magnificent lens. The of zero was chosen for the fatigue testing to ensure crack initiation [38].
color map represents the surface height measured by an SWLI. (For inter- For an initiation notch length of 4.8 mm, the resulting maximum SIF
pretation of the references to colour in this figure legend, the reader is referred was calculated, using the expression given in ASTM E647 for MT
to the web version of this article.) samples, to be 9.83 MPa√m. Prior to applying the load, the sample
surface was imaged using both the 10× and 50× magnification lenses.
initiation location. Measurement using the 10× magnification lens provides a high tem-
poral resolution since the FOV is larger, while measurement using the
50× magnification offers a higher spatial resolution.
2. Experiment method After applying the first cycle, plastic deformations, manifested as
large surface roughness development, were observed near the tips of
2.1. Sample preparation the center notch. A method similar to that described in [38] was used to
measure the size of the plastic zone. The sample surface was imaged
Due to the superior property under harsh mechanical and thermal using the 10× lens before and after applying the first cycle and the
loading, nickel alloys are extensively used in the aerospace industry. difference image between these two images, which accounts for the
Having a face-centric-cubic (FCC) crystal structure, nickel is also known plastic deformation due to the applied load, was constructed. The dif-
to develop noticeable surface topography changes under plastic de- ference image was then segmented into narrow vertical strips so that
formation. The thin oxide layer on the surface of nickel makes the study the surface roughness for each strip can be calculated and plotted
of surface topography changes under fatigue loading possible. versus the distance from the crack tip. The surface roughness steadily
Therefore, annealed nickel with a 99.52% purity was chosen as the decreased with distance from the crack and eventually levered off to a
model material for this study. EBSD inspection of the sample surface value that is comparable to that of the unloaded sample. The distance at
showed a random grain size and orientation, and thus confirmed the which the surface roughness leveled off was identified as the edge of the
behavior of the material to be isotropic. The average grain size mea- plastic zone. For the sample used in this study, the size of the plastic
sured according to ASTM E112-13 planometric procedure was 89 µm. A zone was determined to be 0.4 mm. Accordingly, an area of
static tensile test of a dog-bone sample machined from the as-received 0.8 × 0.8 mm2 (0.1 mm containing the tip of the notch and 0.7 mm in
material determined that the yield stress of the material is 211 MPa. front of the tip) was defined as the region of interest (ROI). As shown in
This value was used to determine the maximum load applied to the Fig. 2, the ROI contains the entire plastic zone area (i.e. the areas in
sample to ensure that the sample, except at the stress concentration blue and cyan) as well as some elastic region (i.e. the areas in yellow
locations, remained elastic throughout the fatigue test. and red) surrounding the plastic zone. In addition, it was observed that
A middle tension (MT) sample designed according to the ASTM E more profound plastic deformations were concentrated in the areas
647 standard, as shown in Fig. 1(a), was machined from the as-received around the two tips of the notch. Therefore, these areas were imaged
nickel sheet. The center notch was machined using Electrical Discharge using the 50× lens to obtain surface topography images with a finer
Machining (EDM) with a 0.254 mm wire, which resulted in a notch resolution. In total, 30 images, each has the size of 234 × 175 µm and a
width of 0.4 mm. After machining, the surfaces of the sample near the predefined overlap of 20% with the neighboring images, were acquired.
tips of the notch were manually polished with conventional sandpapers Using the 50× magnification lens, the resolution of the acquired sur-
of different grit sizes. Afterward, the sample was polished using alu- face height is 1 nm. These images were stitched together using Vi-
mina polish powders down to the size of 1 μm to achieve a mirror-like sion64, resulting in a surface topography image of the entire ROI with
surface. In order to highlight the grain boundaries, the surface was 2186 × 2186 pixels and a lateral resolution of 0.366 μm.
etched using an ASTM E407 – 25 solution, producing a surface The fatigue test was continued up to 1000 cycles with a 10 Hz
roughness of 50 nm. Fig. 1(a) shows the dimensions of the sample after loading frequency. During this stage, the test was paused at every 200
polishing. For reference purpose, the left side of the notch is called the cycles and the sample surface was imaged. Since the surface topography
west notch while the right side of the notch is called the east notch. changes decreased gradually, the fatigue interval was increased to 500
cycles up to 5000 cycles, again increased to 1000 cycles up to 20,000
2.2. Fatigue test procedure cycles, and finally increased to 2000 cycles up to 40,000 cycles. In
addition, SEM images were acquired after the first cycle, at 20,000
To apply cyclic loading, the sample was mounted in a fatigue test cycles, 40,000 cycles, and every 10,000 cycles thereafter. The SEM
bench (BOSE Electroforce) which was capable of applying cyclic images were correlated to the SWLI images by aligning the surface
loading up to 3000 N at a frequency of up to 100 Hz. An SWLI profil- features such as slip bands in the different locations of the sample
ometer was integrated with the fatigue tester to acquire surface topo- surface as well as near the notch. At 40,000 cycles, persistent slip bands
graphy images of the sample surface, as shown in Fig. 1(b). This ar- (PSBs) in the area ahead of the west notch was observed in the SEM
rangement enabled us to image the sample surface without removing it image, indicating crack initiation might occur in the following cycles. In
from the fatigue tester. All surface topography images were acquired order to ensure that we do not miss the crack initiation, the SWLI image

72
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

interval was reduced to 1000 cycles. At 50,000 cycles, the formation of length and has passed four grains on the surface. More than half of the
a micro-crack was confirmed by the SEM image. To observe the crack crack path is perpendicular to the loading direction, either due to a
propagation, the fatigue interval was reduced from 1000 cycles to 500 grain boundary that coincident with this direction or because the mode-
cycles after 60,000 cycles. The experiment was terminated at 95,000 I SIF has become sufficiently large that the crack path is no longer
cycles when the formation of a through-thickness crack was observed. microstructure-sensitive. In addition, a through-thickness crack was
In total, the entire experiment produced 126 SWLI images. The testing observed to have formed on the vertical wall of the notch.
schedule and imaging intervals are shown in Table 1. The corresponding SWLI images showing the surface topography of
the sample are shown in Fig. 4. The slip bands formed just after the first
3. Crack initiation behavior fatigue cycle were clearly visible in the SWLI image shown in Fig. 4(a).
The formation of more slip bands in multiple locations after 20,000
The SEM images of the sample surface near the west notch at dif- cycles similar to the ones observed in the SEM images are shown in
ferent fatigue cycles are shown in Fig. 3. For all images, the loading Fig. 4(b). The surface topography image acquired at 40,000 cycles, as
direction is along the vertical direction. As explained in the previous shown in Fig. 4(c), clearly shows the slip bands in grain 1 and 3.
section, large plastic deformations were observed from the SWLI image However, the slip band in grain 2 where the crack initiated was not
after the first cycle. This plastic deformation zone after the first cycle, clearly visible. This is because the edge of the notch was round off due
however, was not clearly visible in the SEM image shown in Fig. 3(a); to polishing and the SWLI was not able to image this area because the
only very small slip bands are visible. The slip bands were seen to form surface was at a large angle with the image plane. The surface topo-
along a 45 degree angle with respect to the loading direction, indicating graphy image acquired at 50,000 cycles (i.e. when crack initiation was
that the maximum shear stress occurs in this direction. As the test observed from the SEM image) was similar to that acquired at 40,000
continued, more slip bands were created and the height and depth of cycles; however, the surface height of the areas farther from the notch,
the extrusions and intrusions increased. At 20,000 cycles, five locations as highlighted, has decreased; notice that the color has changed from
near the edge of the west notch were observed to have large plastic red in Fig. 4(c) to red-yellow in Fig. 4(d). This surface topography
deformations in the form of slip bands, as shown in Fig. 3(b). These slip changes, as discussed later, are indications of crack initiation. The
bands were mostly formed around the edge of the notch. The size of two surface topography image acquired at 60,000 cycles, shown in Fig. 4(e),
sets of slip bands, i.e. the ones in zone 3 and 4, were bigger than the matches closely with the SEM image shown in Fig. 3(e). Again, it is
other areas. A zoomed-in view of zone 3 and 4 at 40,000 cycles is shown difficult to identify the exact location of the crack tip and there appears
in Fig. 3(c). To determine the grain boundaries, microscopic image to be multiple micro-cracks along the grain boundary between grain 1
using the SWLI profilometer was taken and overlaid on the SEM images. and 4. Similarly, the surface topography image acquired at 90,000
Tracing the slip bands also helped to determine the grain boundaries; cycles, shown in Fig. 4(f), clearly shows the crack patch and matched
the majority of the slip bands were formed in grain 1 and they were closely with the SEM image shown in Fig. 3(f).
stopped near the grain boundary between grains 3 and 4. The number
and size of the slip bands saturated and persistent slip bands (PSBs) 4. Approach
were formed. The slips bands are confined primarily in two grains, i.e.
grains 1 and 3. In grain 1, the slip bands start near the edge of the 4.1. Difference image calculation
notch, lies along the 45 degree direction with respect to the loading
direction, and end at the grain boundary. The slip bands in grain 3 are Since we were interested in quantifying the surface topography
also at a 45 degree angle with the loading direction; they are near a change between fatigue intervals, the images need to be registered to a
triple point (i.e. the intersection of the grain boundaries of grain 1, 2, common coordinate system so that a difference image can be calculated
and 3) and bounded by two grain boundaries. For both grains 1 and 3, by subtracting any two registered images. The surface topography
the slip bands near the triple point had larger intrusions and extrusions, image of the entire ROI acquired after applying the 1st cycle is shown in
both in size and height, than those away from the triple point. Some of Fig. 5(a). The plastic deformations near the notch are clearly visible in
the PSBs may have large inclusions that can be considered as a crack. blue and light green. The boundary of the plastic zone was determined
However, since the SEM image was acquired when the sample was using the method described in [38]. To register the images, the method
unloaded, it was almost impossible to differentiate a crack from large similar to that explained in [38] was employed to find the common
PSBs. Grain 2 did not have significant slip band development as com- features between a reference image and an image of interest. This re-
pared to grain 1 and 3, except a slip band near the edge of the notch. quires converting the surface topography image into a grayscale image
The SEM image around this slip band, acquired at 50,000 cycles, is first by normalizing the height values between zero to one using linear
shown in Fig. 3(d). By tilting the sample, the SEM was able to image the normalization. The grayscale image corresponding to Fig. 5(a) is shown
top surface as well as the vertical wall of the notch. It clearly shows that in Fig. 5(b). The area near the edge of the notch appears in the dark
a crack that extended about 2 μm inside the material had formed. This because the edge of the notch was rounded off due to polishing, as
is the first time that a visible crack was observed from the SEM images, evidenced in the 3D view of the surface shown in Fig. 5(c). To evaluate
indicating that the crack initiated between 40,000 and 50,000 cycles. the effect of this edge rounding on the surface profiling, the surface
The crack length on the surface, however, was difficult to determine
since it was difficult to differentiate a crack from large PSBs based on Table 1
the SEM image. In addition, the exact location of the crack initiation SWLI and SEM imaging intervals (all units in cycle).
cannot be determined; the crack could have initiated near the triple Fatigue cycle Imaging interval
point and propagated toward the edge or initiated at the edge of the
notch. The SEM image of the crack acquired at 60,000 cycles is shown SWLI SEM
in Fig. 3(e). It appears that the crack has propagated along the grain
0 1 1
boundary between grain 1 and 3 and made a left turn at the triple point 1 1 1
between grain 1, 3, and 4. However, it is difficult to pinpoint exactly 1–1000 Every 200 None
where the crack tip was. In addition, there seems to have a few micro- 1000–5000 Every 500 None
cracks at the intersection of the slip bands in grain 1 and the grain 5000–20,000 Every 1000 At 20,000
20,000–40,000 Every 2000 At 40,000
boundary between grain 1 and 4. The crack path becomes much clear in
40,000–60,000 Every 1000 Every 10,000
the SEM image taken at 95,000 cycles, i.e. at the end of the fatigue test, 60,000–95,000 Every 500 Every 10,000
as shown in Fig. 3(f). The crack length has reached about 130 μm in

73
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

Fig. 3. SEM images of the sample surface at different fatigue cycles; (a) after 1 cycle showing faint slip bands; (b) after 20,000 cycles showing five areas with
substantial slip band developments; (c) after 40,000 cycles showing a zoomed-in view of zone 3 and 4; (d) after 50,000 cycles showing the observation of a crack at
the notch wall; (e) after 60,000 cycles showing the bridging of the slip bands at the ends near the grain boundary; (f) after 95,000 cycles showing the crack path.

Fig. 4. Surface topography images of the sample near the notch after: (a) 1 cycle, (b) 20,000 cycles, (c) 40,000 cycles, (d) 50,000 cycles, (e) 60,000 cycles and (f)
95,000 cycles. The color represents the surface height. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

topography image is overlaid with the SEM image and is shown in addition, some pixels in the area just next to the unimaged area had
Fig. 5(d). When using the 50× lens, the SWLI profilometer is capable of erroneous measurements. Therefore, these two areas are removed from
measuring angles up to 26.7 degrees. As a result, the area right next to further analysis. The total width of these two areas was around 7 µm.
the edge of the notch was not imaged because of the rounded edge that Because of this reason, the area near the edge of the notch, which we
resulted in a sharp angle between the surface and the imaging plane. In observed to have a crack propagating into the material at 50,000 cycles,

74
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

Fig. 5. (a) Surface topography image acquired after the 1st cycle. The area shaded in blue shows the plastic zone; (b) grayscale image after 1st cycle showing the
selection of flat area for registration; (c) 3D surface topography image showing the rounded edge of the notch before applying the first cycle; (d) overlaid SWLI image
on SEM image showing un-imaged area due to the rounded edge; (e) registration features identified in the flat area; (f) different image between 0 and 1 cycles
showing out-of-plane deformation due to applied load. The inset highlights the monotonic plastic zone (shaded in blue). (For interpretation of the references to color
in this figure legend, the reader is referred to the web version of this article.)

was not imaged. On the other hand, the area around the triple point, slip bands near the notch, as shown in the zoomed-in images. The
which is most likely to be the crack initiation site, was fully imaged. difference image between these two images is shown in Fig. 6(c). It
The rounded edge also created some difficulties for image registration; represents the cyclic plastic zone, i.e. the change in depth between the
when the rounded area was included for image registration, the de- first cycle and the 20,000 cycles. The zoomed-in view of the cyclic
tected registration points (i.e. the common features such as grain plastic zone indicated that the changes in the slip bands are mostly
boundaries between the reference and the image of interest) were concentrated around the triple point. The difference images at 30,000
concentrated near the notch since the rounded edge has a large height cycles and at 40,000 cycles are shown in Fig. 6(d) and 6(e), respec-
difference from the other areas. Having registration points con- tively. During these two fatigue intervals, the location of the depression
centrating in a small area led to large registration errors. To overcome did not change much but its depth has increased, especially at the area
this problem, an area outside of the plastic zone with an almost flat to the left of the triple point. At 50,000 cycles, however, the depression
surface, as indicated by the rectangular box in Fig. 5(b), was chosen for appears to have moved to the left; the area near the triple point no
image registration. The registration points detected inside this flat area longer have the largest depth and the deepest location has moved at
between the surface topography images taken before and after applying about 50 μm to the left of the triple point, as shown in Fig. 6(f). The
the 1st cycles are shown in Fig. 5(e). The transformation matrix that location change of the depression coincided with the observation of the
accounts for the translation and rotation required for transforming the crack initiation from the SEM image.
second image to the coordinate of the first image was then calculated
from these registration points. By applying the transformation matrix to
4.2. Surface roughness damage indices
the entire image, the registered image was produced. The surface to-
pography changes between the two images can then be calculated by
To quantify the observed surface topography changes, the difference
subtracting the registered images. The difference image between the
images were divided into subsets so that local surface roughness para-
surface topography image before and after applying the first cycle
meters can be calculated. As we have discussed early, the surface to-
which represents the out-of-plane deformation due to the applied load
pography changes are contributed by both the slip bands and the out-of-
is shown in Fig. 5(f). By correcting the rounded edge and the inherent
plane movements of the grains. In order to capture the out-of-plane
initial surface roughness, the different image shows a more symmetric
movements of the grains, the subset size should be sufficiently large. In
surface topography change. The insert in Fig. 5(f) shows the zoom-in
other words, the selection of the subset size is a trade-off between
view of the monotonic plastic zone, i.e. the plastic zone formed after
spatial resolution and prediction reliability; a smaller subset size pro-
applying the first cycle. It manifested as a depression, i.e. an area that
vides a finer spatial resolution but a larger subset size offers a more
has substantially negative height changes.
reliable prediction. Since the grain size varied between 20 and 120 μm,
The surface topography images acquired after the first cycle and the
the subset size was chosen to be 50 × 50 pixels which covers an area of
20,000 cycles are shown in Fig. 6(a) and 6(b). These two images look
18 × 18 μm. The difference image between the 1st and 200 cycles is
almost identical except small changes in the height and formation of
shown in Fig. 7(a) and the surface topography of a subset, capturing the

75
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

Fig. 6. (a) and (b) surface profile after the 1st cycle and 20,000 cycles; (c), (d), (e) and f) difference images between 20,000, 30,000, 40,000, 50,000 cycles and the
1st cycle. These difference images represent plastic deformation development after the 1st cycle, i.e. the cyclic plastic deformation.

surface height variations across grain boundaries, is shown as the inset. cycles was tracked and analyzed for predicting the crack initiation site.
To construct the damage index map, the arithmetic average surface
roughness Sa, as well as the maximum height of the surface St, were 5. Results
calculated for each subset. Since the ROI contained the notch that does
not have a surface, the pixels covering the notch did not have any depth 5.1. Predicting the crack initiation site
value. As a result, these pixels were manually removed from further
analysis. The Sa map calculated from the difference image shown in The difference images between the 1st cycle and the 1,000, 10,000,
Fig. 7(a), after removing the subsets containing the notch, is shown in and 20,000 cycles, respectively, are shown in Fig. 8(a–c) and the cor-
Fig. 7(b). As expected, the subsets near the notch have higher damage responding Sa damage index maps are shown in Fig. 8(d–f). At 1000
indices; three regions, one at around between 10 and 11o’clock, one cycles, the difference image and the corresponding damage index map
near 8o’clock, and the other near 7o’clock appear as “hot-spots”, i.e. are similar to those at 200 cycles (see Fig. 8(a), 8(d) and Fig. 7), i.e. the
localized areas with concentrated large damage indices. Among the difference maps show a small depression formed to the left of the notch
three hot-spots, the hot-spot at 8o’clock has the largest damage indices. and the damage index map shows three “hot spots”. At 10,000 and
The evolution of the damage indices in these areas with the fatigue 20,000 cycles, the depression increased in size as well as in depth, as

Fig. 7. (a) Difference image between 1 and 200 cycles; the inset shows the surface topography of a 50 × 50 pixel subset (b) Sa damage map calculated by dividing a)
into subsets.

76
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

Fig. 8. Difference image between the first cycle and (a) 1000, (b) 10,000, and (c) 20,000 cycles and the corresponding Sa damage index maps after (d) 1000, (e)
10,000, and (f) 20,000 cycles.

shown in Fig. 8(b) and 8(c). However, the Sa damage index became In addition, the three subsets had similar damage indices for the first
more concentrated. At 10,000 cycles, the subsets having large damage few thousand cycles. After 5000 cycles, however, the crack initiation
indices were more concentrated around 8o’clock, as shown in Fig. 8(e). subset started to have a consistently larger damage index increase than
While a few subsets at around 7o’clock still have relatively large da- its two neighboring subsets, even though the damage index increase
mage indices, the area between 10 and 11o’clock can no longer be still has significant fluctuations. A 10-point moving average applied to
considered as a “hot spot”. At 20,000 cycles, one subset at 8o’clock has the damage indices shown in Fig. 10(b) smoothed out the curves sub-
a substantially higher damage index than the other subsets, as shown in stantially, as shown in Fig. 10(c); the crack initiation subset displayed
Fig. 8(f). In comparison, the other two locations had substantially less consistently larger damage increases and a higher increase rate than its
damage index values. At this point, we are confident to predict that the neighbors after 5000 cycles. Based on Fig. 10(c), we can confidently
crack initiated at the subset with the highest damage index. To validate predict the crack initiation site after 10,000 cycles, if not earlier.
this prediction, the damage index map was overlaid on the SEM image The averaged Sa damage index map, constructed from the 10-point
acquired at 20,000 cycles and is shown in Fig. 9. The subset having the moving average of the consecutive height difference images after 1000,
largest damage index coincided with the slip bands that have the largest 10,000 and 20,000 cycles are shown in Fig. 11(a–c). Comparing with
extrusion and intrusion, right next to the triple point. The two subsets the damage index map shown in Fig. 8(b), the averaged damage index
just above and below the subset having the largest damage index, map shown in Fig. 11(b) clearly shows that the crack initiation subset
marked as subset 1 and 2, also had relatively large damage indices. had the largest damage index. In other words, the crack initiation site
The variations of the Sa damage indices of the crack initiation subset can be predicted from the averaged Sa damage index map at 10,000
(i.e. the subset having the largest Sa at 20,000 cycles) and its two im-
mediate neighbors are compared in Fig. 10(a). While the crack initia-
tion subset has a consistently larger Sa than its neighbors shortly after
15,000 cycles, the differences are relatively small. Besides, the damage
indices for all three subsets fluctuated despite the overall increasing
trend; neighboring subset 1 had significant fluctuations with the da-
mage index exceeded that of the crack initiation subset at around
10,000 cycles. This fluctuation is due to the fact that the damage indices
were calculated from the difference image between a specific fatigue
cycle and the first cycle. As such, they only account for the accumula-
tive surface topography changes between these two cycles. On the other
hand, the fluctuation of the damage index indicated that some of the
surface topography changes are reversible (i.e. strain or height ratch-
eting). In order to account for the height ratcheting, the difference
images were calculated between consecutive fatigue intervals and the
resulting Sa damage indices for the three subsets are plotted in
Fig. 10(b). The Sa damage indices increase were relatively large for the
first few fatigue intervals, consistent with the observations that early
fatigue intervals generated relatively large plastic deformations. After
the initial increase, however, the damage indices increase reduced to a Fig. 9. Damage index map overlaid on the SEM image; both acquired after
lower level, indicating that the material hardens with the fatigue cycle. 20,000 cycles. The subset with the maximum damage index coincided with the
triple point area experiencing large slip band developments.

77
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

a 140 b 140 c 100


Crack Initiation subset Crack Initiation subset 90 Crack Initiation subset
120 Subset 1 120 Subset 1 Subset 1
Subset 2 Subset 2 80
100 100 Subset 2
70
60

Sa (nm)

Sa (nm)
Sa (nm)

80 80
50
60 60 40
40 40 30
20
20 20
10
0 0 0
0 5000 10000 15000 20000 0 5000 10000 15000 20000 0 5000 10000 15000 20000
Fatigue cycle Fatigue cycle Fatigue cycle

Fig. 10. The evolution of the Sa damage index with increasing fatigue cycles for three subsets in the critical area; (a) Sa damage indices calculated using the image
after the first cycle as the reference; (b) Sa damage indices calculated using the image acquired at previous fatigue interval as the reference; (c) Sa damage indices
averaged over 10 fatigue intervals.

cycles, as compared to at 20,000 cycles based on the damage index map initiation site, as shown in Fig. 12(a). The Sa damage indices of these
calculated from the difference image using the image at first cycle as five subsets are plotted versus the fatigue cycles in Fig. 12(b). For subset
the reference. A similar trend was observed for the compound St da- 1, the Sa damage index increased steadily up to 42,000 cycles and re-
mage index, as shown in Fig. 11(d–f). Since St is calculated from the mained steady afterwards. This behavior can be explained by ex-
absolute difference between the maximum and minimum height, it is amining the local stresses before and after crack initiation. Before crack
more susceptible to the measurement or registration errors, as evi- initiation, subset 1 was likely to have large local stresses since it was at
denced by the number of isolated pixels having relatively larger St than the crack initiation site. When the crack initiated, a crack tip was
their neighbors. Nevertheless, the crack initiation site can be still de- generated at a location in front of subset 1. In other words, subset 1 is
tected just after 10,000 cycles. now behind the crack tip and no longer has large stresses. Due to this
stress reduction/release, subset 1 experiences very little additional
plastic deformation after crack initiation. As a result, the damage index
5.2. Detecting crack initiation time
of subset 1 stabilized after crack initiation. The stabilization of the
damage index implies that the crack might have initiated around
In addition to predicting the crack initiation site, tracking the var-
42,000 cycles, which is again consistent with the SEM observation. The
iation of the damage index with the fatigue cycle may also enable us to
same trend was observed for subset 2 except that the damage indices
determine the time at which the crack initiated; it is expected the da-
decreased slightly after the crack initiation. This decrease can be related
mage index will increase steadily before crack initiation and stabilize
to the elastic energy release at crack initiation. After the decrease, the
after the crack initiation. To validate this hypothesis, five subsets were
damage indices increased until approximately 55,000 cycles when the
selected for investigation: subset 1 corresponds to the crack initiation
crack reached this subset, after which the damage index remained
subset, subset 2 is the subset immediately to the left of subset 1, and
constant. The damage indices of subsets 3–5, i.e. the subsets that are
subset 3–5 are three random subsets that were away from the crack

Fig. 11. Moving averaged Sa damage index maps after (a) 1 cycle, (b) 10,000 cycles. and (c) 20,000 cycles; and moving averaged St damage index maps after (d) 1
cycle, (e)10,000 cycles, and (f) 20,000 cycles.

78
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

300.0
b Subset 1 Crack Initiation time
250.0 Subset 2
Subset 3
200.0 Subset 4
Subset 5

Sa (nm)
150.0

100.0

50.0

0.0
0 20000 40000 60000
Fatigue cycle

Fig. 12. (a) Location of subsets at which the variations of the damage indices were studied; (b) variation of Sa damage index versus fatigue cycles at different
locations.

away from the crack initiation site, increased steadily with the fatigue of the registration errors, etc. Two modifications were made in the
cycle. However, their magnitudes were considerably lower than those present work and are worth discussing. In addition, a few interesting
of subset 1 and 2, indicating the surface topography changes are small observations pertinent to crack initiation are also worth discussing.
at these subsets. Registration error: In this study, due to the rounded edge of the
To correlate the damage index variations to the surface topography notch, the area farther from the notch, i.e. the elastic region outside of
changes, the change of the surface profile, along a vertical line passing the plastic zone, was chosen for image registration, which could in-
the crack initiation subset (see Fig. 13(a)), was investigated. The sur- troduce some registration errors due to the in-plane plastic deforma-
face profile changes were calculated with respect to the surface profile tions. However, it expected that the plastic deformations are pre-
after applying the first cycle. The surface profile contained some dominately along the out-of-plane direction and the in-plane plastic
“noise”, i.e. small height changes contributed by the slip lines or re- deformations are relatively small. Moreover, it was observed that the
gistration errors. To remove the “noise”, a Butterworth Low Pass Filter majority of the plastic deformation occurred at the first cycle, and the
(LPF) was applied and the filtered height changes at fatigue intervals of plastic deformation was much smaller in subsequent cycles. Using the
5,000 cycles are shown in Fig. 13(b). The curve at 20,000 cycles clearly image acquired after the first cycle as the reference, registration using
shows a depression spanning from around 350 μm to 450 μm. The depth the elastic region may not introduce significant registration errors to
of this depression increased with the fatigue cycling up to 40,000 cy- the plastic region. To assess the amount of registration errors, ten
cles. At 45,000 cycles, however, the depth of the depression rebounded, characteristic geometric features, such as the corners of the grain
becoming considerably less than the previous cycles. This surface re- boundaries, in the areas close to the notch were identified manually and
bound indicates that the crack may have initiated between 40,000 to the errors of their respective registered locations were calculated to be
45,000 cycles and this is consistent with the SEM observation as well as 1.3 pixels in average. This value is comparable to the registration error
the prediction of the Sa damage index. The surface profile changes, reported in [40]. In other words, using the elastic region for image
therefore, can provide a quantified index for detecting the time at registration did not seem to introduce additional registration errors.
which the crack initiated. Calculation of compound damage index: A modification is made to
how the fluctuation of the damage index due to strain ratcheting is
accounted for; [40] introduced a compound damage index based on a
6. Discussion weight averaging formula while the present work smoothens the da-
mage index curve using a 10-point moving average. The moving
The image registration technique and the formulation of the surface average was introduced based on the variation of the damage index
roughness damage index presented in this work are similar to those shown in Fig. 10; it is simpler to calculate and more straightforward to
described in [40], in which the damage index was used to predict the explain.
future propagation path of a microstructure-sensitive crack. The readers Crack initiation from the sample surface versus from the bore of the
should refer to [40] for more detailed discussions on the selection of the notch: crack initiation in a polycrystalline sample depends on two
subset size, the temporal resolution of the SWLI imaging, the influence

500.0
b
0.0
Height (nm)

-500.0

20,000 cycles
-1,000.0
26,000 cycles
30,000 cycles
-1,500.0 36,000 cycles
40,000 cycles
45,000 cycles
-2,000.0
300.0 350.0 400.0 450.0 500.0 550.0
Y (μm)

Fig. 13. (a) Location of the vertical line along which the surface profile changes were investigated; the vertical line is parallel to the loading direction; (b) surface
profile change at different fatigue cycles along the defined line.

79
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

factors, namely the stress state and the weakest grain. For thin samples, steadily with fatigue cycling before the crack initiation but remained
the stress state on the surface is plane stress while the state of stress in steady after the crack initiation. Therefore, the crack initiation time can
the bore is plane strain. Generally, plane stress is more critical for crack be determined based on the saturation of the damage index. The crack
initiation [42]. Previous studies have shown that, when dominated by initiation time can also be determined by tracking the depth of the
the plane stress condition, cracks primarily initiated from the edge of depression near the crack initiation site; it increased steadily before
the notch [43,44] because the grains on the edge of the notch are less crack initiation and rebounded after the crack has initiated. This study
constrained and deform easier [45]. This is the reason that most crack established the surface roughness parameter as a reliable damage index
initiation studies were focused on the surface of thin samples for crack initiation site prediction as well as for crack initiation detec-
[13,22,24,25,32,36,37,46,47]. Another consideration is that crack in- tion. Since this study is purely empirical, the presented technique could
itiation starts from the grains that are favorably oriented for plastic be combined with EBSD, X-ray tomography, 3D DIC strain measure-
deformation. When the thickness increases and more grains exist in the ment techniques and crystal plasticity simulation in the future to gain a
bore of the notch, the possibility that the weakest grain (i.e. the fa- mechanistic understanding of crack initiation mechanisms.
vorably oriented grain) exists in the bore increases. In this case, the
crack may initiate in the bore of the notch instead of the surface. It Acknowledgments
would be difficult to measure the surface topography of the bore surface
using an SWLI. However, the surface will develop roughness as soon as This work has been funded by the Air Force Office of Scientific
the crack-tip plastic zone reaches the surface. Therefore, it is still pos- Research (AFOSR) grant FA9550-14-1-0319.
sible to predict at which location the crack will break the surface based
on the surface roughness damage index of the surface. Appendix A. Supplementary material
Applying the surface roughness damage index for studying other
crack initiation mechanisms: This study focused on the crack initiation Supplementary data to this article can be found online at https://
behavior of a pristine sample in a benign environment. In this case, the doi.org/10.1016/j.ijfatigue.2019.02.027.
surface topography change is a result of the out-of-plane plastic de-
formation due to the microstructure heterogeneity in a polycrystalline References
sample. Since the damage indices defined in this study are calculated
directly from the surface topography changes, they provide a quanti- [1] Boeff M, Hassan HU, Hartmaier A. Micromechanical modeling of fatigue crack in-
tative characterization of plastic deformation that is the precursor for itiation in polycrystals. J Mater Res 2017;32(23):4375–86.
[2] Suresh S. Fatigue of materials. 2nd ed. Cambridge, UK: Cambridge University Press;
crack initiation. This technique is applicable for the study of other crack 1998. p. 704.
initiation mechanisms as long as the material experiences plastic de- [3] Zhu L, Wu ZR, Hu XT, Song YD. Investigation of small fatigue crack initiation and
formation before crack initiation. For example, corrosive environments growth behaviour of nickel base superalloy GH4169 Available from Fatigue Fract
Eng Mater Struct [Internet] 2016;39(9):1150–60http://doi.wiley.com/10.1111/ffe.
will generally make the material more susceptible to plastic deforma- 12430.
tion [48,49]. Similar to crack initiation in the air, crack usually forms in [4] Krupp U. Fatigue Crack Propagation in Metals and Alloys: Microstructural Aspects
the surface in corrosive environments [50]. Therefore, the present and Modelling Concepts; 2007.
[5] Raabe D, Sachtleber M, Zhao Z, Roters F, Zaefferer S. Micromechanical and mac-
technique could potentially be used to study crack initiation under romechanical effects in grain scale polycrystal plasticity experimentation and si-
different corrosive environments. mulation. Acta Mater 2001;49(17):3433–41.
[6] Hochhalter JD, Littlewood DJ, Christ RJ, Veilleux MG, Bozek JE, Ingraffea AR, et al.
A geometric approach to modeling microstructurally small fatigue crack formation:
7. Conclusion
II. Physically based modeling of microstructure-dependent slip localization and
actuation of the crack nucleation mechanism in AA 7075–T651. Model Simul Mater
The surface topography changes of a pure nickel sample with a Sci Eng 2010;18(4).
center notch were investigated under fatigue loading and quantitatively [7] Wan VVC, MacLachlan DW, Dunne FPE. Integrated experiment and modelling of
microstructurally-sensitive crack growth. Int J Fatigue 2016;91:110–23.
analyzed to predict the crack initiation site as well as to determine the [8] Sangid MD. The physics of fatigue crack initiation. Int J Fatigue 2013;57:58–72.
crack initiation time. We observed that https://doi.org/10.1016/j.ijfatigue.2012.10.009.
[9] Mughrabi H. Cyclic slip irreversibilities and the evolution of fatigue damage. Metall
Mater Trans B Process Metall Mater Process Sci 2009;40(4):431–53.
1) Plastic deformation led to the formation of a depression near the [10] Mughrabi H. Cyclic slip irreversibilities and the evolution of fatigue damage. Metall
ends of the notch after applying the first cycle. The depression Mater Trans B 2009 Apr;40(4):431–53.
deepened with fatigue cycling and moved forward after crack in- [11] Mughrabi H, Wang R, Differt K, Essmann U. Fatigue crack initiation by cyclic slip
irreversibilities in high-cycle fatigue 5-5–41 Fatigue Mech Adv Quant Meas Phys
itiation; Damage [Internet]1983.
2) Slip bands were formed at multiple locations near the ends of the [12] Manonukul A, Dunne FPE. High- and low-cycle fatigue crack initiation using
notch after applying the first cycle. The slip bands increased in polycrystal plasticity. 2004;460(2047):1881–903.
[13] Risbet M, Feaugas X. Some comments about fatigue crack initiation in relation to
height as the fatigue cycle increases. One of these slip band systems cyclic slip irreversibility. Eng Fract Mech. 2008;75(11):3511–9.
eventually led to crack initiation between 40,000 and 50,000 cycles, [14] Wan VVC, Maclachlan DW, Dunne FPE. A stored energy criterion for fatigue crack
as confirmed by SEM imaging. nucleation in polycrystals. Int J Fatigue 2014;68:90–102.
[15] Yeratapally SR, Glavicic MG, Hardy M, Sangid MD. Microstructure based fatigue life
prediction framework for polycrystalline nickel-base superalloys with emphasis on
Based on these observations, the surface roughness parameters were the role played by twin boundaries in crack initiation. Acta Mater
introduced as the fatigue damage index and evaluated for crack in- 2016;107:152–67.
itiation site prediction. The surface roughness parameters were calcu- [16] Sangid MD, Maier HJ, Sehitoglu H. A physically based fatigue model for prediction
of crack initiation from persistent slip bands in polycrystals. Acta Mater
lated from the surface topography changes between a particular fatigue 2011;59(1):328–41.
cycle and a reference image. Two different types of reference images, [17] Man J, Obrtlík K, Polák J. Extrusions and intrusions in fatigued metals. Part 1. State
i.e. the surface topography images acquired after the first cycle and the of the art and history. Philos Mag 2009;89(16):1295–336.
[18] Efthymiadis P, Pinna C, Yates J, Street M, Sheffield S. Fatigue Crack Initiation in
one acquired at the previous fatigue interval, were studied. When using AA2024: a coupled in-situ mechanical testing and crystal plasticity study. Fatigue
the surface topography image acquired after the first cycle as the re- Fract Eng Mater Struct 2018.
ference image, the crack initiation site was predicted at 20,000 cycles. [19] Abuzaid WZ, Sangid MD, Carroll JD, Sehitoglu H, Lambros J. Slip transfer and
plastic strain accumulation across grain boundaries in Hastelloy X. J Mech Phys
When the damage index was calculated from the surface topography Solids 2012;60(6):1201–20.
changes between consecutive fatigue intervals and smoothened using a [20] Carroll J, Abuzaid W, Lambros J, Sehitoglu H. An experimental methodology to
10-point averaging, it predicted the crack initiation site at 10,000 cy- relate local strain to microstructural texture. Rev Sci Instrum. 2010;81(8):1–9.
[21] Risbet M, Feaugas X, Guillemer-Neel C, Clavel M. Use of atomic force microscopy to
cles. In addition, the damage index at the crack initiation site increased

80
J. Fathi Sola, et al. International Journal of Fatigue 124 (2019) 70–81

quantify slip irreversibility in a nickel-base superalloy. Scr Mater 1.OE.55.10.104108.


2003;49(6):533–8. [35] Vladimirov AP, Kamantsev IS, Veselova VE, Gorkunov ES, Gladkovskii SV. Use of
[22] Ho HS, Risbet M, Feaugas X. On the unified view of the contribution of plastic strain dynamic speckle interferometry for contactless diagnostics of fatigue crack initia-
to cyclic crack initiation: impact of the progressive transformation of shear bands to tion and determining its growth rate Available from Tech Phys [Internet]
persistent slip bands. Acta Mater 2015;85:155–67. 2016;61(4):563–8http://link.springer.com/10.1134/S106378421604023X.
[23] Cretegny L, Saxena A. AFM characterization of the evolution of surface deformation [36] Jha DK, Singh DS, Gupta S, Ray A. Fractal analysis of crack initiation in poly-
during fatigue in polycrystalline copper. Acta Mater 2001;49(18):3755–65. crystalline alloys using surface interferometry. EPL 2012;98(4):3–8.
[24] Risbet M, Feaugas X, Guillemer-Neel C, Clavel M. Damage in nickel base superalloy: [37] Haghshenas Ali, Khonsari MM. Damage accumulation and crack initiation detection
influence of local parameters measured by electron backscattered diffraction and based on the evolution of surface roughness parameters. Int J Fatigue
atomic force microscopy. Scr Mater 2009;60(5):269–72. 2018;107:130–44. https://doi.org/10.1016/j.ijfatigue.2017.10.009.
[25] Ho HS, Risbet M, Feaugas X, Bigerelle M, Zhang E. Surface roughness based char- [38] Kelton R, Fathi Sola J, Meletis EI, Huang H. Visualization and quantitative analysis
acterization of slip band for damage initiation in a nickel base superalloy. of crack-tip plastic zone in pure nickel. JOM 2018;70(7):1175–81. https://doi.org/
Medziagotyra 2018;24(1):112–7. 10.1007/s11837-018-2865-5.
[26] Polák J, Man J. Initiation of stage i fatigue cracks – Experiments and models. [39] Kelton R, Fathi Sola J, Meletis EI, Huang H. Study of the surface roughness evo-
Procedia Eng. 2015;101(C):386–94. lution of pinned fatigue cracks, and its relation to crack pinning duration and crack
[27] Kamaya M. Observation of fatigue crack initiation and growth in stainless Steel to propagation rate between pinning points. 2017;1–5.
quantify low-cycle fatigue damage for plant maintenance. E-Journal Adv Maint [40] Fathi Sola J, Kelton R, Meletis EI, Huang H. A surface roughness based damage
2013;5:185–200. index for predicting future propagation path of microstructure-sensitive crack in
[28] Kimura H, Akiniwa Y, Tanaka K, Kondo J, Ishikawa T. Observation of Fatigue Crack pure nickel. Int J Fatigue 2019;122:164–72. https://doi.org/10.1016/j.ijfatigue.
Initiation and Early Propagation in Ultrafine-Grained Steel By Atomic Force 2019.01.012.
Microscopy. 10th Int Congr Fract [Internet]. 2001;ICF100877OR. Available from: [41] Petzing J, Coupland J, Leach R. The Measurement of Rough Surface Topography
http://www.gruppofrattura.it/ocs/index.php/ICF/ICF10/paper/viewFile/4869/ using Coherence; 2010.
6876. [42] Jack AR, Price AT. Effects of thickness on fatigue crack initiation and growth in
[29] Vehoff H, Nykyforchyn A, Metz R. Fatigue crack nucleation at interfaces. Mater Sci notched mild steel specimens. Acta Metall 1972;20(7):857–66.
Eng A 2004;387–389(1–2 SPEC. ISS.):546–51. [43] Sova JA, Crews JH, Exton RJ. Fatigue-crack initiation and growth in notched 2024-
[30] Jiang R, Karpasitis N, Gao N, Reed PAS. Effects of microstructures on fatigue crack T3 specimens monitored by a video tape system. NASA Tech NOTE. 1976;1(August
initiation and short crack propagation at room temperature in an advanced disc 1976).
superalloy. Mater Sci Eng A 2015;641:148–59. https://doi.org/10.1016/j.msea. [44] Leis BN, Galliher RD. Growth of physically short corner cracks at circular notches.
2015.05.065. Am Soc Test Mater 1982:399–421.
[31] Nakai Y. Evaluation of fatigue damage and fatigue crack initiation process by means [45] Wu Z, Sun X. Multiple fatigue crack initiation, coalescence and growth in blunt
of Atomic-Force Microscopy. Mater Sci Res Int 2001;7(2):1–9. notched specimens. Eng Fract Mech 1998;59(3):353–9.
[32] Stoudt MR, Levine LE, Creuziger A, Hubbard JB. The fundamental relationships [46] Stoudt M, Hubbard J. Analysis of deformation-induced surface morphologies in
between grain orientation, deformation-induced surface roughness and strain lo- steel sheet. Acta Mater 2005 Sep;53(16):4293–304.
calization in an aluminum alloy. Mater Sci Eng A 2011;530:107–16. https://doi. [47] Stoudt MR, Hubbard JB, Leigh SD. On the relationship between deformation-in-
org/10.1016/j.msea.2011.09.050. duced surface roughness and plastic strain in AA5052 - Is it really linear? Metall
[33] Wang Y, Meletis EI, Huang H. Quantitative study of surface roughness evolution Mater Trans A Phys Metall Mater Sci 2011;42(9):2668–79.
during low-cycle fatigue of 316L stainless steel using Scanning Whitelight [48] Asphahani AI, Sridhar N. Corrosion fatigue of nickel and nickel-base alloys.
Interferometric (SWLI) Microscopy. Int J Fatigue 2013;48:280–8. https://doi.org/ Corrosion 1982;82:587–95.
10.1016/j.ijfatigue.2012.11.009. [49] Srivatsan S, Sudarshan TS. Mechanisms of fatigue crack initiation in metals: role of
[34] Muravsky LI, Picart P, Kmet AB, Voronyak TI, Ostash OP, Stasyshyn IV. Evaluation aqueous environments. J Mater Sci. 1988;23(5):1521–33.
of fatigue process zone dimensions in notched specimens by two-step phase shifting [50] El May M, Saintier N, Palin-Luc T, Devos O, Brucelle O. Modelling of corrosion
interferometry technique. Opt Eng [Internet]. 2016;55(10):104108 Available from: fatigue crack initiation on martensitic stainless steel in high cycle fatigue regime.
http://opticalengineering.spiedigitallibrary.org/article.aspx?doi=10.1117/ Corros Sci 2018;133:397–405. https://doi.org/10.1016/j.corsci.2018.01.034.

81

You might also like