Core microstructure-dependent bending fatigue behavior and crack growth
Core microstructure-dependent bending fatigue behavior and crack growth
Keywords: Carburizing is a thermo-chemical surface treatment through which a very hard martensitic layer develops in the
Carburizing external surface (case) of steel components resulting in substantial improvement in the fatigue life. Nevertheless,
Microstructure the overall fatigue properties of carburized steel components are yet severely dependent on the microstructure
Bending fatigue which develops in the interior region (core). This paper deals with the effects of core microstructure on bending
Fatigue crack growth
fatigue behavior and fatigue crack growth of carburized steel parts. V-notched steel specimens were fabricated
Hardness
and subjected to two case hardening cycles where, respectively, bainitic-martensitic and ferritic-bainitic-mar-
tensitic microstructures developed in the core regions supported by similar fully martensitic microstructures in
the case. 4-point plane bending fatigue tests were conducted to study the fatigue behavior of the heat-treated
specimens. Furthermore, the effects of the core microstructures on fatigue crack growth resistance were also
investigated. Hardness measurements revealed that both batches of specimens have similar hardness properties
on the exterior surfaces, in the case-hardened layers and also in the cores. Moreover, the results showed that the
specimens with the bainitic-martensitic core microstructure provide a marginally better fatigue performance in
the finite life regime as compared to the ferrite-containing counterparts. More noticeable difference was,
however, observed in the corresponding endurance limits where the former demonstrated a higher magnitude
than the latter. Besides, the bainitic-martensitic core microstructure resisted the fatigue crack propagation more
effectively than the ferrite-containing specimens.
*
Corresponding author.
E-mail address: [email protected] (H. Farivar).
https://doi.org/10.1016/j.msea.2019.138040
Received 14 April 2019; Received in revised form 9 June 2019; Accepted 15 June 2019
Available online 15 June 2019
0921-5093/ © 2019 Elsevier B.V. All rights reserved.
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
the constitutive phases and the subsequent effects on overall fatigue components on the plane bending fatigue performance and fatigue
behavior. crack propagation which have rarely been reported in the literature.
Several factors influence the fatigue performance of case-hardened Through performing two heat treatment cycles, different multi-phase
steel components [6]. For instance, component's geometry and size [7], microstructures develop in the core while hardness properties of the
loading condition [8], applied heat treatment and the resulting mi- case regions are retained similar. Hence, the core microstructure-de-
crostructure [9–14], surface hardness [15], case hardness depth pendent fatigue behavior and crack growth can solely be investigated.
[16,17], surface oxidation, surface compressive residual stresses
[18–20], surface roughness [21–23], amount, size and distribution of
non-metallic inclusions [24–26] and environmental temperature [27] 2. Material and methods
and moisture [28] are the main influential factors. In the work of Jo
et al. [29] the fatigue behavior of a carburized Cr–Mo steel under cyclic 2.1. Material and specimen
stress conditions including axial, rotating bending, torsional and in-
phase axial-torsional fatigue tests were investigated. It was shown that The chemical composition of the case hardening steel investigated in
the long life (at 106 cycles) fatigue strength under axial load was about this study is indicated in Table 1. The amounts of the alloying elements are
30% lower than bending fatigue strength resulting from the difference in the range of the common case hardening steel 20MoCr4 which is widely
in stress gradient between the two types of tests. Based on the obtained employed in automotive industry for fabrication of gear shafts and gear
experimental results, a simple two-layer model was presented taking wheels. The material was laboratory melted in a vacuum induction fur-
the case and core material properties into account which could accu- nace and casted in an 80 kg ingot with a dimension of
rately predict the axial cyclic deformation behavior for the carburized 140 × 140 × 500 mm3. The ingot was further homogenized at 1200 °C for
steel. Zhang et al. [9] investigated the fatigue crack growth behavior of 3 h and underwent multi-step hot forging process producing two smaller
a high-speed locomotive axle steel processed by induction hardening blocks each with a final dimension of 70 × 70 × 1000 mm3. The hot-
that contains a hardened surface layer with gradient microstructure. It forged blocks were thereafter left in still air and cooled down to room
was shown that fatigue crack growth rate decelerated first and then temperature. It was then followed by normalizing process at 900 °C to
accelerated with the increase of crack length within the gradient layer. remove the effects of prior forging process and acquire a homogeneous
It was further discussed that the process of crack growth deceleration microstructure. Standard Charpy V-notch specimens (CVN,
and arrest is mainly resulted by the gradient microstructure with large 10 × 10 × 55 mm3, DIN EN ISO 148-1:2011-01) were fabricated out of
compressive residual stresses. In another work [30], the effects of car- the normalized blocks along the longitudinal direction. The V-notched
burizing and micro-defects on competing failure behaviors of Ni–Cr–W specimens were later on employed for conducting the bending fatigue
steel under gigacycle fatigue using axial loading method was in- tests. The V-shaped notch generates an intensive localized stress field at
vestigated. The authors concluded that the interior failure induced from the notch root which ensures that during the bending fatigue tests, crack
small inclusion or microstructural inhomogeneity is the predominant initiates and propagates in a certain area of the specimens enabling an
failure mode in the life regime beyond 105 cycles. It was further argued easier analysis of the crack growth. The linear elastic stress concentration
that the case carburizing can only improve the fatigue strength with factor (Kt) was calculated to be 3.69 according to the finite element si-
surface failure and has no influence on the fatigue strength with interior mulations performed in software Abaqus (Static/General) with a mesh size
failure. Dengo and coworkers [31] studied the bending fatigue behavior of 0.1 mm and C3D8I elements.
of MnCr and NiCrMo case-hardened steels employing plane bending
fatigue tests on plain and notched specimens. Based on the obtained Table 1
experimental data, it was suggested that irrespective of the load type Chemical composition of the investigated steel (wt%).
(either bending or axial), a plateau at the endurance limit exists. C Si Mn P S Cr Mo Al N Fe
Moreover, it was found that the residual stresses in the interior of the
specimens increased due to the applied fatigue loads which was at- 0.18 0.39 0.75 0.010 0.022 0.35 0.50 0.03 0.0069 Bal.
tributed to the transformation of retained austenite to martensite,
known as TRIP effect. The effectiveness of retained austenite on im-
proving the bending fatigue lifetime of carburized steel parts has also 2.2. Heat treatments
been reported elsewhere [32,33]. In the research conducted by Yu et al.
[34] the strong relationship between carbide morphologies and me- The heat treatment cycles applied on the V-notched specimens are
chanical properties in carburized steels were demonstrated. The uni- schematically shown in Fig. 1. The process of carburizing and the fol-
formly distributed fine spherical carbides (with a Fe3C structure) im- lowing high-pressure gas (N2) quenching were conducted in a dual-
proved the tensile and bending fatigue properties due to the chamber vacuum carburizing furnace. The fabricated specimens were in-
suppression of crack propagation and initiation, respectively. It was itially heated above the Ac3 temperature1 followed by the vacuum car-
further shown that networked carbides along grain boundaries, how- burizing at 950 °C using Acetylene (C2H2). Afterwards, the temperature
ever, facilitate the crack nucleation and propagation and hence dete- was reduced at a rate of 4 °C/min to the hardening temperature of 860 °C
riorate the tensile and bending fatigue properties. In another work [35], (cycle A) and held for 15 min. The specimens finally subjected to high-
the effects of microalloying element Nb in different contents on rotating pressure (14 bar) gas quenching to reach room temperature. However, in
bending fatigue properties of three case hardening steels were in- another charge (cycle B), upon completion of the carburizing step, the
vestigated. It was reported that the case-hardened steels alloyed with specimens were cooled down to a hardening temperature lower than that
0.04 wt% Nb have larger number of fine Nb (C, N) precipitates showing conducted in cycle A, namely 775 °C, followed by an identical holding
the best fatigue performance as compared with the Nb-free and the time. Since 775 °C is lower than the Ar3 temperature2 of the base material
0.08 wt% Nb-alloyed counterparts. The obtained results were explained (i.e.: core), hence, austenite in the core partly decomposes into ferrite. This
by the prior austenite grains refining effect of Nb and the associated batch of specimens was likewise quenched to room temperature. The
effective role in controlling the fatigue resistance. Similar observation
on the relationship between prior austenite grain size and fatigue 1
Ac3: 865 °C, determined by dilatometry, is the critical temperature above
properties was also reported in the paper published by Liu et al. [36] which a single phase of austenite is solely stable.
where the rotating bending fatigue limit of carburized steels containing 2
Ar3: 800 °C, determined by dilatometry, is the critical temperature at which
fine AlN precipitates were considerably improved. ferrite starts to transform out of the parent austenite during the cooling at the
This paper studies the role of core microstructure in carburized steel rate of 4 °C/min.
2
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 1. Schematic illustration of the case hardening cycles applied on V-notched specimens consisting of vacuum carburizing, isothermal hardening at 860 °C (cycle
A) and 775 °C (cycle B) followed by high-pressure gas (N2) quenching and the final tempering step.
3
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Table 2
Quantitative fractions of the core microstructural phases (point-counting
method).
Specimen ferrite bainite martensite
C860_15 – 86 ± 5% 14 ± 5%
C775_15 30 ± 2% 23 ± 3% 47 ± 2%
Fig. 5. Light optical micrographs of the case-hardened V-notch specimens in different regions, namely in core (a–b) and at the notch root (c–d).
4
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 6. EBSD results obtained from the core section of the case-hardened V-notch specimens. Image quality (IQ), kernel average misorientations (KAM) and retained
austenite (RA) maps are respectively shown. The KAM values were calculated with a step size of 100 nm and over a radius of 300 nm with a maximum misorientation
of 5°. Note that the indexed volume fractions of retained austenite (RA, i.e.: the dispersed constituents marked in black) is presented in the lower right corner of the
corresponding figures, (c) and (f).
Fig. 7. (a) Carbon distribution profiles at the notch root of the carburized specimens measured by EPMA method. (b) The corresponding microhardness profiles
measured at the notch root. (c) Light optical micrographs show the impressions left by microhardness indentations.
front of the ground and polished surface of the loaded specimens taking and W are the specimen's thickness and width, respectively, as per the
high-resolution images in 5-s intervals. The captured images were terminology adopted in standards ASTM E647 [38] and ASTM E399
subsequently evaluated using the image processing software ARAMIS®.
The fatigue crack propagation experiments were carried out under the
( )
[39]. a is the crack length and f W is a geometry function which reads
a
as follows:
same conditions as described in section 2.4.1.
Besides, Eq. (1) was used to calculate the stress-intensity factor
range, ΔK.
a a
1.99 ( )(1
a
W
a
W ) (2.15 3.93 ( ) + 2.7 ( ) )
a
W
a 2
W
f =3 3
W W
2 (1 + 2 )(1 )
a a 2
W W
PS a
K= 3
f (2)
BW 2 W (1)
It is to be mentioned that Eq. (1) is a slightly modified form of the
where ΔP is the algebraic difference between the maximum and original equation addressed in standard ASTM E399 which is routinely
minimum applied forces in a cycle, S is the equivalent span length, B employed for calculation of ΔK in standard SENB specimens under
5
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 8. Carbon distribution maps measured in the core microstructures of specimens C860_15 and C775_15 are in (a) and (b), respectively. (c–e) SE images of the
areas where the line scan measurements were carried out. Note the longitudinal contamination area left by the measurements. The corresponding carbon con-
centrations across the lines are in (f).
Core hardness [HV10] 286 ± 4 300 ± 3 Fig. 4 displays cross-sectional views of the case-hardened V-notched
Surface hardness [HV30] 735 ± 17 743 ± 23 specimens showing the microstructures developed across the speci-
mens’ width. From the presented light optical images, it is observed that
at the outer surface (case) of the heat-treated specimens both, a mar-
quasi-static loadings. The external load, P, in the original equation,
tensitic layer with similar penetration depth was developed. The mi-
which is in the form of static state, was replaced by ΔP in the current
crostructures appeared in the interior section (core) are different,
investigations converting the equation to a dynamic state. Span length,
though.
S, was also re-calculated from a 3-point bending testing condition into
The microstructural features in the core and notch root of the case-
the current 4-point bending state. The equivalent S equals to 4 L , where
3
hardened specimens can be seen in Fig. 5 with more details. As it is
L is the distance between the lower support rollers, 40 mm (Fig. 2).
shown in the light optical micrographs, the core microstructure of
Moreover, Paris’ law, Eq. (3), was used to analyze the fatigue crack
specimens C860_15 is comprised of bainite and martensite, Fig. 5.a.
growth rate during the steady propagation of the fatigue crack.
However, the core microstructure of specimens C775_15, in addition to
da bainite and martensite, also consists of ferrite, Fig. 5.b. The develop-
= C Km ment of ferrite is caused by performing the isothermal treatment at the
dN (3)
reduced hardening temperature where ferrite is thermodynamically
where a and ΔK, as already mentioned above, are the fatigue crack stable. The fractions of the microstructural phases developed in the
length and stress-intensity factor range, respectively, N is the number of cores were quantified and listed in Table 2. As pointed out above, at the
cycles, hence, da is the fatigue crack growth rate, C and m are fitting notch root of the specimens both, a fully martensitic microstructure
dN
constants. with similar depths is observed, Fig. 5c-d. The formation of the mar-
tensitic layers is due to the high level of carbon enrichment resulted
6
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 9. S–N diagrams of the case-hardened V-notched specimens obtained in 4-point cyclic plane bending tests. Se represents the corresponding endurance limit for
each test series.
from the carburizing process. By moving further from the case towards region where the corresponding base carbon content (0.18 wt%) is
the core, bainite gradually appears next to the martensite in the mi- reached. It is to be noted that the scattering in carbon content observed
crostructures of both specimens. This region, which is known as tran- in the core region (deeper than 600 μm) is attributed to the unequal
sition area, is less enriched in carbon as compared to the case, thus, partitioning of carbon atoms between different phases developed in the
giving rise to a lower chemical stability of austenite and its decom- core. The corresponding microhardness profiles are also presented in
position into bainite during the quenching step, Fig. 5c-d. It is also Fig. 7.b, which are also similar with some minor differences. According
worth mentioning that the average prior austenite grain size (PAGS) for to the obtained microhardness profiles, the thickness of the hardened
both batches of specimens was calculated to be approximately 20 μm layers (CHD3) below the notch root in specimens C860_15 and C775_15
which is due to the similar austenitization and carburization conditions. is measured to be approximately 415 μm and 450 μm, respectively.
The core microstructures of the investigated specimens were further Light optical images taken from the notch root of the hardened speci-
analyzed by EBSD method and the obtained results are presented in mens show the impressions left by microhardness indentations, Fig. 7.c.
Fig. 6. According to the image quality (IQ) maps, the core micro- In Fig. 8 the distribution of carbon atoms among the microstructural
structure of specimens C860_15 is mostly bainite, but there is also a constituents developed in the core regions are presented. Fig. 8.a and
minor fraction of martensite present, Fig. 6.a. However, large areas Fig. 8.b show the carbon distribution maps measured in the specimens
with high image quality can be detected within the IQ map of speci- C860_15 and C775_15, respectively. As it is readily seen, there are many
mens C775_15 implying the presence of ferrite in addition to bainite film-like and globular regions in the core microstructure of specimens
and martensite, Fig. 6.d. The ferritic areas can also be easily observed in C860_15 which are significantly enriched with carbon implying the
kernel average misorientation (KAM) map of specimens C775_15 which presence of retained austenite (RA). The carbon-rich constituents are
corresponds to the regions with lower KAM values, Fig. 6.e. Unlike mostly located in the bainitic phase. Due to the formation of bainitic
specimens C775_15, no ferrite was detected in KAM map of specimens ferrite (corresponds to the carbon-depleted regions) and the associated
C860_15, Fig. 6.b. In Fig. 6.c and Fig. 6.f the fraction, morphology and limited carbon solubility, the excess carbon atoms diffuse out into the
distribution of retained austenite (RA) in specimens C860_15 and adjacent austenitic phase. This, therefore, increases the chemical sta-
C775_15 can be observed, respectively. In specimens C860_15, ap- bility of the carbon-enriched austenitic regions a part of which is re-
proximately 3.6% RA was indexed. The measured fraction of RA in mained untransformed after quenching to room temperature [42].
specimen C775_15 is however only 0.3%. It is worth noting that both Since the fraction of bainite in the core microstructure of specimens
typical morphologies of RA, known as film-like [40] and globular [41] C775_15 is much lower than that of specimens C860_15 (Table 2),
are present in the developed core microstructures. hence, the fraction of carbon-enriched regions and consequently the
fraction of retained austenite are accordingly lower. These findings are
3.2. Carbon distributions and hardness properties in accordance with the quantified fractions of retained austenite mea-
sured by the EBSD analyses (Fig. 6). The results of the carbon line scans
Fig. 7.a demonstrates the carbon distribution profiles at the notch
root of the carburized specimens. It is easily observed that the carbon
profiles of both the specimens are similar in the carburized region with 3
Case hardness depth, abbreviated as CHD, is routinely used as an index in
an approximate carbon content of 0.6 wt% at the surface. The carbon case-hardened specimens and corresponds to the depth with a hardness value of
concentration gradually decreases by moving further towards the core 550 HV.
7
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 10. Example images of the in-situ fatigue crack growth analyses at stress
amplitude 1370 MPa taken by the camera during the cyclic bending tests on
specimens C860_15 and C775_15 over different cycles. The images show the
growth of fatigue cracks initiated at the notch root propagating towards the
core region.
8
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 11. Experimental results showing the variations of the (a) fatigue crack length versus number of cycles, and (b) fatigue crack growth rate versus stress intensity
factor range for the investigated specimens at stress amplitude 1370 MPa.
at around 7500 cycles in specimens C775_15. This is however about transformation-induced plasticity (TRIP) effect alleviating the stress
10000 cycles on average for specimens C860_15. Moreover, the fatigue field ahead of the crack tip [11,32,43,44] is an additional factor
crack in the former propagates considerably faster than that in the latter yielding to better fatigue properties as compared to specimens C775_15.
and can grow no longer than approximately 2.6 mm prior to the failure.
In test series C860_15, the fatigue crack can however grow up to ap- 3.5. Fractographic analysis
proximately 4 mm before the specimens fail. More detailed data on the
effect of core microstructure on fatigue crack growth resistance are Cross-sections of the fracture surfaces were analyzed to study the
presented in Fig. 11.b. As it is obviously seen, specimens C775_15 have damaged area in the core microstructures caused by the intensive stress
greater values of da/dN (fatigue crack growth rate) in comparison to field near the fatigue crack tip during its propagation. The example LOM
specimens C860_15 implying a faster fatigue crack propagation in the and SEM images of the cross-sectional analyses carried out for the spe-
core microstructure of the former. Whereas, the core microstructure of cimens loaded at stress amplitude 1370 MPa are shown in Fig. 12. From
the latter can better withstand against the fatigue crack propagation the LOM images, Fig. 12.a and Fig. 12.b, which give a broad overview of
having lower da/dN values. Furthermore, the respective exponents m, the cross-sectioned surfaces, it is readily seen that the core micro-
which denote the slopes of the presented curves in the Paris regime structure of specimen C775_15 is deformed more severely than that of
(marked with dashed lines on the curves), are also shown in Fig. 11.b. specimen C860_15 and also to a larger extent. In the presented SEM
The Paris exponent m varies between 5.6-5.8 and 1.9–2.2 for specimens micrographs, Fig. 12.c and Fig. 12.d, the lateral extent of plastic zones
C775_15 and C860_15, respectively, clearly reflecting a more sluggish along with the fatigue-induced microstructural deformations are seen
fatigue crack growth in the latter as compared to the former. The cal- with more details. Unlike specimen C775_15, the core microstructure of
culated constants m and C for the investigated specimens are sum- specimen C860_15 was only slightly deformed. Besides, the traces of
marized in Table 4. plastic deformation in the former can be identified much farther to the
crack propagation path, in a zone with an approximate width of 270 μm.
Table 4 Whereas, the deformed area in the latter was barely wider than 100 μm.
Calculated constants of Paris’ equation for the investigated specimens. The SEM image of specimen C775_15, Fig. 12.d, reveals that most of the
Specimen C860_15_01 C860_15_02 C775_15_01 C775_15_02 identified secondary cracks are located on the boundary of ferrite and
martensite. This is attributed to the sharp difference in local strength of
C 2.6234 E−8 5.7678 E−8 5.7605 E−13 3.5660 E−13 ferrite and martensite resulting in severe strain discontinuity and stress
m 2.2 1.9 5.6 5.8
localization at the boundary of the constituents during the loading.
Therefore, the interfaces of ferrite and martensite act as preferred sites
for crack formation [45,46] when the core microstructure of specimen
C775_15 subjects to the intensive stress field generated near the tip of the
According to the obtained results, the better fatigue performance of growing fatigue crack. Furthermore, due to the severe strain partitioning,
specimens C860_15 can be attributed to the bainitic core microstructure the soft phase of ferrite undergoes plastic deformation giving rise to
with the associated high-angle grain boundaries deflecting the fatigue formation of the fatigue striations in this phase [47]. The secondary
crack during the cyclic loadings [12]. The presence of retained auste- cracks and fatigue striations are indicated by the dashed circles and ar-
nite in the core microstructure of this test series and the resulting rows in the presented SEM images, respectively.
9
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 12. Cross-sectional views of the near-fracture zones in the specimens loaded at stress amplitude 1370 MPa analyzed by (a) light optical microscopy, and (b)
scanning electron microscopy, showing the fatigue-induced secondary cracks and the fatigue striations next to the fatigue crack propagation path.
Additionally, the fracture surfaces of the V-notched specimens were constituting phases causes a severe strain incompatibility at the interfaces
investigated from a top view as well revealing the corresponding morpho- of ferrite-bainite and ferrite-martensite leading to large stress concentra-
logical features in the case and core regions. Example optical stereo- tions at these boundary sites. This results in premature localized plastic
micrographs and SEM images of the fracture surfaces of the specimens deformation in the weaker phase (ferrite) at the vicinity of the interfaces
loaded at stress amplitude 1370 MPa are presented in Fig. 13. As it is ob- before the global yield strength of the specimens is reached. Furthermore,
vious from Fig. 13.a, the fatigue crack propagation zone and the rupture this gives rise to debonding of the interfaces which is constantly promoted
zone can easily be identified in specimen C860_15. These regions in the over the course of the cyclic loading creating micro-voids and micro-cracks
fracture surface of specimen C775_15, however, cannot be differentiated, at these zones, as seen in Fig. 12.d. During further cycling, the fatigue-
Fig. 13.b. Besides, further analyses show that the fracture mechanism in the induced micro-voids and micro-cracks grow, the ligaments between them
case regions of both the specimens are primarily governed by a brittle mode are necked and eventually teared (micro-voids coalescence) which leads to
which is due to the hardened martensitic layer with the high carbon con- final failure of the specimens. This, correlates well with the fracture surface
centration resulted from the carburizing process. The corresponding SEM of these specimens with a ductile appearance where many dimples were
images, Fig. 13.c and Fig. 13.d, clearly show the intergranular facets im- detected, Fig. 13.f. Therefore, the generation of micro-voids and micro-
plying a brittle fracture. The fracture analysis in the core region of specimen cracks at the ferrite-bainite and ferrite-martensite interfaces due to the
C860_15, Fig. 13.e, shows the characteristics of quasi-cleavage fracture large strain incompatibility and sharp stress concentration decreases the
which is attributed to the presence of bainite and martensite constituents in work hardening capability of the specimens and thus, degrades the overall
the core microstructure of the mentioned specimens. By contrast, the fatigue performance of the ferrite-containing specimens as compared to
fracture surface in the core region of specimen C775_15 was mainly covered specimens C860_15.
by dimples, Fig. 13.f, implying that in this region ductile fracture me- Moreover, it is generally argued that the high-cycle fatigue behavior of
chanism is dominant. This can be explained by the presence of ferrite grains high-strength steels are primarily governed by crack nucleation which is, in
next to bainite and martensite in the core of specimens C775_15. turn, initiated by high local stress concentration. Unlike specimens
Furthermore, it is well known that ferrite is the softest (i.e.: weakest) C775_15, the core microstructure of specimens C860_15 is believed to
constituent in multi-phase steel alloys. In specimens C775_15, the core cause a much lower strain incompatibility with substantially less localized
microstructure is composed of considerable fraction of ferrite beside of stress concentration. This is basically due to the existence of the micro-
bainite and martensite which are much harder phases. When these speci- structural constituents with a lower strength difference (bainite and mar-
mens subject to cyclic loading, the significant difference in strength of the tensite) in the latter leading to lower stress concentration at the
10
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
Fig. 13. Optical stereomicrographs of the fracture surfaces are in (a, b). SEM micrographs taken from the case and core of the failed specimens are in (c, d) and (e, f),
respectively. Typical fatigue-induced micro cracks and fatigue striations are indicated by arrows in the SEM images of the core regions. The presented images were
taken from the fracture surface of the specimens loaded at stress amplitude 1370 MPa.
corresponding phase boundaries. This, in turn, leads to a higher slip band (C860_15) show a slightly better bending fatigue performance in the
initiation stress level in specimens C860_15 and only when the applied finite life regime than the ferrite-containing counterparts (C775_15).
stress is high enough can trigger the fatigue crack nucleation, hence, a The endurance limit of specimens C860_15 and C775_15 lays at
higher endurance limit for specimens C860_15 in comparison to specimens 747 MPa and 685 MPa, respectively.
C775_15. 3. Fatigue crack growth analyses revealed that the bainitic-martensitic
core microstructures are more resistant against the crack propaga-
4. Conclusions tion as compared to the ferrite-containing microstructures. This is
attributed to the fatigue crack-deflecting effect of bainitic micro-
V-notched steel specimens subjected to different case hardening cycles structures associated with numerous high-angle grain boundaries.
and subsequently underwent 4-point plane bending fatigue experiments. Additionally, the presence of retained austenite in the bainitic-
Fatigue crack growth tests were additionally performed to examine the martensitic core microstructure assist to alleviate the stress level
fatigue crack resistance of the developed core microstructures. The con- ahead of the crack tip (TRIP effect), hence, slows down the fatigue
clusions derived from this study can be summarized as follows: crack growth rate.
4. The interfaces of ferrite and martensite in the core microstructure of
1. By applying the designed case hardening cycles, distinct multi-phase specimens C775_15 act as preferred sites for crack formation during
microstructures develop in the core of the specimens, whereas, fully the cyclic loadings. The sharp difference in local strength of the
martensitic microstructures appear in the case regions having si- constituting phases results in severe stress localization on the
milar hardness properties. The core hardness properties in both the boundaries. This leads to a more extended plastic zone in vicinity of
specimens are also similar. the fatigue crack propagation path in specimens C775_15 than
2. The specimens with a bainitic-martensitic core microstructure specimens C860_15.
11
H. Farivar, et al. Materials Science & Engineering A 762 (2019) 138040
12