J Engfailanal 2019 104166

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Engineering Failure Analysis 106 (2019) 104166

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Strain-controlled fatigue and fracture of AISI 410 stainless steel


T
⁎ ⁎
Muhammed J. Adinoyi , Nesar Merah , Jafar Albinmousa
Department of Mechanical Engineering, King Fahd University of Petroleum and Minerals, Dhahran, Saudi Arabia

A R T IC LE I N F O ABS TRA CT

Keywords: Microstructural characteristics, monotonic and strain-controlled cyclic axial behaviors of AISI
Microstructure 410 stainless steel were investigated. Grain size and distribution is uniform throughout the alloy's
Cyclic softening microstructure. Tensile behavior is characterized with high ductility yet mild strain hardening.
Fatigue life Strain-controlled fatigue testing was conducted at strain amplitudes ranging from 0.3% to 0.9%.
Fractography
Cyclic stress softening was observed at all applied strain amplitudes. Hysteresis loops show that
Ductile fracture
stress-strain stabilization was not reached by the alloy. Cyclic softening rate decreases with fa-
tigue life and its ratio was independent of strain amplitude above 0.5%. An accumulating plastic
damage behavior was exhibited by the alloy. Fatigue life data were well correlated with Coffin-
Manson and plastic energy density damage equations. Multiple cracks were observed in speci-
mens under monotonic and cyclic loadings.

1. Introduction

Steels remain one of the mostly used structural alloys because of their toughness and availability in a variety of grades. Stainless
steel grades are primarily designed to resist corrosion and have applications in a wide range of components [1–5]. In addition, they
can withstand high temperature applications [6,7]. Among the commonly used stainless steel are the austenitic-based 316 and 304
grades. The austenite steel has been extensively investigated with varying conclusions on the strain-stress behavior and the fatigue
life. Yuan et al. [8] observed a rapid increase in stress followed by gradual cyclic softening for 316LN stainless steel tested under
strain-controlled loading. For a cast duplex steel, Chen et al. [9] found that there was initial hardening in the stress-cycle followed by
secondary hardening, attributed to martensitic phase transformation in the austenite. The damage mechanism in stainless steel is a
critical factor in quantifying the fatigue performance. It is reported by Kamaya [10] that both surface crack and bulk damage consist
of internal and localized cracks that influence the fatigue life of 316 stainless steel. The author [10] observed that cyclic stress
evolution show an initial rapid hardening followed by gradual increase in stress. Golanski and Mrozinski [11] studied the low-cycle
behavior of martensitic cast steel and they reported cyclic softening for all levels of investigated strain amplitudes. The underlying
reason for the different behavior under strain-controlled loading for similar grades of steel can be attributed to the substructural
change during cycling. Therefore, damaging mechanisms vary from one steel to another. Equally important among the stainless steels
is AISI 410 whose application spans a wide range of engineering fields. AISI 410 steel has good corrosion resistance and is used in the
construction of plastic molds, screws for extruders, valves, shafts, bearings and blades in compressors and turbines [12–15]. In spite
of its application in high strained components, where localized plastic deformation can result in fatigue failure, there is yet a dearth of
information about its fatigue and fracture behavior. The most recent failure investigation on AISI 410 is by Moreira et al. [16] who
studied a failed piston rod in spillway floodgate showing that the failure emanated from improper heat treatment during manu-
facturing, leading to high δ-ferrite content, temper embrittlement, low toughness, and susceptibility to stress corrosion cracking.


Corresponding authors.
E-mail addresses: jadinoyi@kfupm.edu.sa (M.J. Adinoyi), nesar@kfupm.edu.sa (N. Merah), binmousa@kfupm.edu.sa (J. Albinmousa).

https://doi.org/10.1016/j.engfailanal.2019.104166
Received 21 February 2019; Received in revised form 27 July 2019; Accepted 28 August 2019
Available online 31 August 2019
1350-6307/ © 2019 Elsevier Ltd. All rights reserved.
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

Table 1
Material chemical composition.
Element Cr C Fe P Mn Si S

wt% 12.1 0.11 86.7 0.018 0.7 0.33 0.002

Hence, the aim of this study is to investigate the strained-controlled fatigue characteristic of AISI 410 alloy. Microstructural analysis
and monotonic testing of the alloy will also be carried out to determine its grain structure and quasi-static behavior. The results from
this investigation will serve as the first-principle fatigue analysis for the material in applications where fatigue failure is probable for
a component subjected to a variety of strain loads.

2. Material and experimentations

The AISI 410 stainless steel investigated in the present study is a 25.4 mm-diameter bar stock of a-meter length. The chemical
composition of the material is illustrated in Table 1. Following ASTM E3 [17],
Microstructure analysis was conducted on specimens taken from the locations, T1 and T2, in the transverse orientation of the steel
bar and from E1 and E2 along its length (extrusion orientation), as annotated in Fig. 1. In both orientations, subscript 1 denotes
samples cut from the edges, while subscript 2 indicates that sample was taken from the center of the stock. Samples were ground,
polished with diamond paste and etched in Nital solution before microscopic observation. Tensile test specimen, with a gauge length
of 25.0 mm and a gauge diameter of 6.0 mm, (Fig. 2(a)), was machined parallel to the extrusion direction. Quasi-static tensile tests
were conducted at a crosshead speed of 2 mm/min, using Instron 5569 tensile test machine, according to ASTM E8-08 Standard [18].
Extensometer with 12.5 mm gauge was used to measure strains up 1%, which is the range of strain required for fatigue tests, it was
then removed and strain above 1% estimated from crosshead displacement. Three tensile tests were performed and the resulting force
and strain data, automatically recorded in the computer during the test, were analyzed to determine the tensile properties of the
alloy.
The strain-controlled axial fatigue tests were conducted as per ASTM E606-12 Standard [19] on smooth solid specimens
(Fig. 2(b)) using Instron servo-hydraulic test frame with a load capacity of 100 kN. The solid specimens were prepared on a CNC
machine with aerospace certified coolant. Gauge sections were machined in one pass for a uniform profile. This was followed by
three-stage polishing with aluminum oxide abrasive belts of grades P240 (59 μm), P400 (39 μm) and P800 (22 μm), respectively.
Alignment cells were used on test fixture to minimize off-axis load and incidence of buckling. The extensometer used has a gauge
length of 11 mm. Test strain amplitudes were varied between 0.3% and 0.9%. Following testing, fractographic analysis was con-
ducted on fractured specimens of selected strain amplitudes under JOEL JSM6610 Scanning Electron Microscope.

3. Results and discussion

3.1. Microstructure

Fig. 3 shows the optical metallographs of the microstructure of AISI 410 stainless steel alloy from the representative locations in
both the transverse and extrusion orientations. The microstructure is a tempered martensite structure with grain size of about 5.0 μm,
calculated by line intercept method. No variation is observed both in the structure and grain sizes of the alloy in both the extrusion
and transverse orientations. The darker areas in the microstructure are attributed to the presence of interlath carbide within mar-
tensite [12]. Martensite structure results from rapid quenching from austenite phase and it is responsible for high strength of
martensitic steels. Similar structure has been reported by Krishna and Bandyopadhyay [14] for a hardened and tempered AISI 410
steel.

3.2. Quasi-static tensile behavior

A typical engineering stress-strain curve for the alloy is presented in Fig. 4. The curve possesses both elastic and plastic regions
with the alloy achieving an average ultimate stress of 591 MPa and a yield strength of 489 MPa. The characteristics curve indicates an

Fig. 1. Annotation of the metallographic procedure.

2
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

Fig. 2. (a) Solid specimen for monotonic tensile test. (b) Solid specimen for axial strain-controlled fatigue test.

Fig. 3. Optical micrograph of etched AISI 410 stainless steel in transverse and extrusion directions.

initial monotonic hardening, followed by rapid softening after yielding. The strain-hardening coefficient, K, and exponent, n, are
found by performing a regression curve fitting on the stress-plastic strain data, resulting in a fitting curve represented by Eq. (1).

σ = 551.0 ε 0.05 (1)

3
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

700

600

Engineering tennsile stress (MPa)


500

400

300

200

100

0
0 5 10 15 20 25
Engineering tensile strain (%)

Fig. 4. Engineering stress-strain curve.

The average mechanical properties, obtained from three tensile tests, are summarized in Table 2. The tensile strength compares to
the value reported for similar alloy by Hejripour and Aidun [13]. The strength of AISI 410 can be linked to its tempered martensite
microstructure previously shown in Fig. 3. Tempered martensite is a ferrite-cementite phase with enhanced ductility and toughness
arising from the ferrite. The cementite phase acts as reinforcement for the ferrite matrix accounting for the relatively high strength
[20]. In addition, the small size of the grains means that many grain boundaries are present within a unit volume. These boundaries
act as barriers to dislocation motion making it difficult for dislocations to be transmitted during plastic deformation, thereby giving
the alloy its relatively high strength. However, as can be seen from Fig. 4 and Table 2, low strain work hardening exponent is
recorded for the alloy, indicating that only slight strain hardening occurs post yielding.

3.3. Low cycle fatigue behavior of AISI 410

3.3.1. Cyclic stress-strain evolution and softening behavior


Fig. 5 shows the variation of the maximum and minimum stresses with the number of applied cycles for the present AISI 410 alloy
and for all the applied strain amplitudes, εa. It can be inferred from this figure that the alloy cyclically softens to failure at all applied
strain amplitudes. The continuously decreasing stress indicates that stress stabilization was not reached by the alloy. It is obvious
from Fig. 5 that the softening behavior is independent of applied strain amplitudes, as the cyclic stress curves at all strain amplitudes
converge to a single path. Softening behavior is usually anticipated for high strength alloys [21]. Ductile and initially strong alloys
soften when subjected to cyclic loading because of the rearrangement of dislocations, which offer less or no resistance to dislocation
mobility. The extent of softening otherwise known as softening ratio (SR) can be estimated from Eq. (2), where σmax and σHNf are
maximum stress and stress at half-life, respectively.
σmax − σHNf σHNf
SR = =1−
σmax σmax (2)

Fig. 6 illustrates the variation of the softening ratio (SR) with the applied strain amplitudes. The SR only rises with increasing
strain amplitude until about 0.45%. From 0.5% onward, the SR evens out and becomes independent of the applied strain. This
phenomenon may be explained by the fact that at strain amplitudes above 0.4%, there is no increase in the ratio of dislocation density
to dislocation mobility. A seemingly slight strain hardening in cyclic stress with the number of cycles for the first few cycles can be
observed at lower strain amplitude of 0.3% and 0.4%. The duration of strengthening is longer for lower applied strain amplitude. This
initial rise in the stress is an elastic characteristics attributable to the stiffness of the material since this happened only at strain

Table 2
Summary quasi-static tensile properties.
Engineering tensile property AISI 410

Tensile strength (MPa), Uts 591 ± 10


0.2% Yield stress (MPa) 489 ± 23.8
Elastic Modulus (GPa), E 208 ± 0.21
Strain hardening coefficient, K (MPa) 551 ± 13.93
Strain hardening exponent, n 0.05 ± 0.02
Reduction in area (%) 61 ± 0.01
Elongation (%) 3.6 ± 0.3
Fracture stress (MPa) 340 ± 7.04
Fracture strain (%) 21 ± 0.98

4
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

800
0.30% 0.40% 0.50% 0.60% 0.70% 0.80%

Maximum & Minimum Stresses (MPa)


400

-400

-800
1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06
Number of fatigue cycles

Fig. 5. Cyclic stress curves for AISI 410 tested at different strain amplitudes.

0.3
Softening ratio

0.2

0.1
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Applied strain amplitude (%)

Fig. 6. Softening ratio versus the applied strain amplitude.

amplitudes substantially within the elastic limit.


The trend of cyclic stress with fatigue life ratio is shown in Fig. 7. Fatigue life ratio at a specified strain amplitude is the ratio of
fatigue life (Ni) at a given time to the number of cycles at failure (Nf) for the specified strain amplitude. A rapid decrease in stress is
observed initially, followed by a steadily decreasing stress, and a final rapid softening which culminates in failure. Similar ob-
servation is reported by Branco et al. [22] for high strength martensitic steel and by Mazánová et al. [23] for 316 L stainless steel. The

600
0.30% 0.40% 0.50%

0.60% 0.70% 0.80%

500
Stress amplitude (MPa)

400

300

200
0 0.2 0.4 0.6 0.8 1 1.2

Fatigue life ratio, Ni/Nf

Fig. 7. Cyclic stress amplitude vs. fatigue life ratio for AISI 410.

5
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

continuous decrease in stress can be linked to the rearrangement of dislocations as earlier observed while the abrupt change at Ni/Nf
above 0.8 is mainly due to damage by rapid crack propagation.
Fig. 8 illustrates the evolution of hysteresis loops of AISI 410 stainless steel alloy at strain amplitudes of 0.3%, 0.6%, 0.7% and
0.9%. The loops are chosen to show the evolution of the hysteresis shape and size as both the strain and number of cycles are
increased. Peak stress drops with the number of cycles while the width of the hysteresis loop increases with the number of cycles. The
former indicates continuous cyclic softening while the latter signifies accumulation of fatigue damage due to the continuous increase
in plastic strain. The trend shown by the cyclic stress evolution in Fig. 5, where there is an initial cyclic hardening at lower strain
amplitude for the early few cycles, then followed by cyclic softening is also revealed in the hysteresis loops evolution. Furthermore,
the evolution of hysteresis loops show that softening decreases with the number of cycles. However, in the latter cycles, the loops
appear to be superimposed on each other, seemingly concealing the drop in peak stress from cycle to cycle that represent cyclic
softening. Accordingly, stress stabilization was not achieved; rather a decreasing cyclic softening rate describes the hysteresis loop
evolution of the present alloy. The hysteresis loops are found to be symmetrical, signifying that the predominant deformation mode is
by slip mechanism. Additionally, serration-like features are present towards the end of the compression reversal at strain amplitudes
higher than 0.6%. This can be attributed to dynamic strain ageing which can occur for both ferrous and non-ferrous alloy due to
annealing, temperature, strain amplitude and rate [24–30]. Dynamic strain ageing (DSA) is a concept of increased flow stress owing
to mobile solid atoms repeatedly locking moving dislocation [24]. Stress increases because dislocations attempt to break loose from
locking by the solute atom or precipitates [31]. It can be observed that the serrations appear at strain amplitude greater than 0.6%
and only in the unloading branch of the loop. This is unlike for most alloys in which serrations manifest in both the loading and
loading branches of hysteresis loop. Therefore, this can suggest that the phenomenon of DSA in the present alloy may be different
from similar other alloys and it is most likely sensitive to strain amplitude and load direction.
Plastic strain, εp, can be measured from hysteresis loop where the stress is theoretically zero or where the loop crosses the stress
axis. More often, it is convenient to evaluate it using Eq. (3).
σa
εp = εa −
E (3)
where εa and σa are, respectively, applied strain and the mid-life stress amplitudes, and E is the modulus of elasticity determined from
the quasi-static tensile test. The relationship between the plastic strain amplitude, εp and stress amplitude σa is expressed in Eq. (4):

σa = K ′ (εp )n′ (4)


where K′ and n′ are the fatigue strength coefficient and exponent, respectively. Fig. 9 shows that the half-life stress amplitude
increases continuously with plastic strain amplitude. Following regression analysis on log-log curve, the cyclic deformation constants
were determined and the plastic strain-stress equation for the present material can be described by Eq. (5):
σa = 881.5(εp )0.124 (5)

3.3.2. Fatigue life equations


The fatigue life is the number of cycles to failure for a given strain amplitude. Failure is defined in this study as the number of
cycles at which load drops to 50% of its maximum value during the fatigue cycle. The relationship between the strain amplitudes and
the fatigue life in strain-controlled fatigue can be represented by:
σ ′f
εe = (2Nf )b
E (6)

εp = ε′f (2Nf )c (7)

σ ′f
εa = εe + εp = (2Nf )b + ε′f (2Nf )c
E (8)
where εe, εp and εa are the elastic, plastic and total strain amplitudes, respectively. While σf′ and b are elastic fatigue constants known
as the strength coefficient and exponent respectively, εf′ and c are the ductility coefficient and exponent, respectively. Eq. (8) is the
Coffin-Manson representation of the curve fitting the strain-life data.
The variation of number of reversals (twice the fatigue life, Nf) with strain amplitudes is illustrated in Fig. 10. The elastic and
plastic strains correlations with fatigue life are also included in the same figure. For the range of strain amplitudes investigated, the
majority of the fatigue life is in the intermediate fatigue life region. The effect of strain amplitude on the specimen life is well
represented by the Coffin-Manson, Eq. (9). The transition life, Nt, below which the plastic strain is the controlling strain was found to
be around 10,000 cycles. As can be observed from Fig. 10, at fatigue life above transition life, Nt, the strain-life curve approaches the
elastic line, while at fatigue life below Nt, the strain-life curve inclines towards the plastic strain line. Consequently, slender hysteresis
loops as typified by low strain amplitude in Fig. 8(a) are present rightward of the transition life. On the other hand, bloated hysteresis
loops (Fig. 8(b)–(d)) characterizing substantial plastic deformation are located leftward of the transition life.
705.08
εa = (2Nf )−0.058 + 0.1565(2Nf )−0.445
208000 (9)
Plastic strain is considered as being essential for damage in a strain-controlled test. With a bulging hysteresis loop as both the

6
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

2nd cycle 5th cycle 50th cycle


1/4 life cycle 1/2 life cycle 3/4 cycle life
600

(a) a= 0.3%

200

Axial stress (MPa)


-200

-600
-0 .4 -0 .2 0 0. 2 0.4
Axial strain amplitude (%)

2nd cycle 5th cycle 50th cycle


1/4 life cycle 1/2 life cycle 3/4 life cycle
800
(b) a= 0.6%

400
Axisal stress (MPa)

-400

-800
-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8
Axial strain amplitude (%)

2nd cycle 5th cycle 50th cycle


1/4 life cycle 1/2 life cycle 3/4 life cycle
800

(c) a= 0.7%
400
Stress (MPa)

Serrations

-400

-800
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1

Strain (%)

2nd cycle 5th cycle 50th cycle


1/4 life cycle 1/2 life cycle 3/4 life cycle
800

(d) a= 0.9%
400
Stress (MPa)

Serrations
-400

-800
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
Strain (%)

(caption on next page)


7
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

Fig. 8. Hysteresis loop evolution for AISI 410 stainless steel for selected axial strain amplitudes studied.

1000

Stress amplitude (MPa)

100
1. 0E-0 3 1 .0 E - 0 2
Plastic strain amplitude

Fig. 9. Cyclic stress-strain curve.

0.1
Exp. total strain
Exp. elsatic strain
Axial strain amplitude (mm/mm)

Exp. plastic strain

0.01 Coffin-Manson line

0.001

0.0001
1.0E+02 1.0E+03 1.0E+04 1.0E+05
Number of reversals to failure, 2Nf

Fig. 10. Axial strain-life curves for the AISI 410.

applied strain and number of cycles increase, it is instructive to quantify damage by evaluating the plastic energy per unit volume.
Plastic energy density, ΔWp, can be evaluated experimentally from hysteresis loops or from the cyclic behavior parameters related by
Eq. (10). Where σa and εp are as defined for Eq. (3), and n′ is the exponent in Eq. (4).

1 − n′ ⎞
∆Wp = 4σa εp ⎛
⎝ 1 + n′ ⎠ (10)

Fig. 11 shows that there exists a good agreement between the energy density evaluated from the area of the hysteresis loops and
that predicted from Eq. (10). This is an indication that the plastic energy density for this alloy can be easily derived from the cyclic
stress-strain parameters in close agreement with that from the hysteresis loop. Fig. 12 illustrates the variation of the plastic energy
density with the number of fatigue cycles for low, medium and high strain amplitudes. Because of the observed initial strain
hardening, the energy density exhibits continuous rise at low strain amplitudes. However, at medium and high strain amplitudes
where plastic strains have fully developed, after few cycles, the energy density remains almost constant. Accordingly, the correlation
between plastic energy, at half-life, and the number of reversals at failure is presented in Fig. 13. The fitting curve relating plastic
energy density with the number of reversals can be represented with Eq. (11).

∆Wp = 554(2Nf )−0.559 (11)

The material constants with the numerical values of 0.559 and 554 are the plastic energy-based fatigue damage exponent and
coefficient, respectively.

3.3.3. Fractography
The fracture mode for the alloy in monotonic tension is illustrated in Fig. 14. The gross decrease in cross section area (Fig. 14(a))
gave rise to the considerably high percentage reduction in area presented in Table 2. Fractured surface is dull with serrated edges,
which is a characteristic of ductile alloys. The SEM fractograph shown in Fig. 14(b), reveals an irregular surface with cracks

8
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

experimental data prediction line

12

Predicted energy density (MJm-3)


10

0
0 2 4 6 8 10 12

Experimental energy density (MJm-3)

Fig. 11. Predicted energy density vs. Experimental energy density.

100
0.30% 0.40% 0.60%
Plastic energy density (MJm-3)

0.70% 0.90%

10

0.1
1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05
Number of fatigue cycles

Fig. 12. Plastic energy density vs. the number of fatigue cycles.

Eq. (10) Exp


100
Plastic energy density (MJm-3)

10

0.1
1.0E+02 1.0E+03 1.0E+04 1.0E+05
Number of reversals to failure

Fig. 13. Plastic energy density versus the number of reversals to failure.

emanating from the edge towards the center of the specimen. The magnified view of the central portion of the fractured surface,
shown in Fig. 14(c), reveals the presence of numerous micro cracks. The development of multiple cracks is an indication that several
slips are responsible for the specimen failure.
Fig. 15 shows the global view of fractured samples tested under fatigue at the strain amplitudes of 0.3, 0.7 and 0.8%. The crack
initiation sites and slow crack propagation areas are marked as (I). The fast crack growth (or unstable crack growth) and the final
fracture zones are labeled (II) and (III), respectively. As can be seen from the annotation with the red arrows, multiple crack initiation

9
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

Fig. 14. Fractography of monotonic tensile specimen.

sites are present, irrespective of applied strain. Similar observation was made by Sakar et al. [32], for sintered stainless steel. The
different areas of crack growth and propagation were measured revealing that the lower the strain, the longer the initiation and slow
crack propagation life. The texture of the various areas of the fatigue phases is different due to the different phases of fatigue events.
Magnifications of the different fracture regions are shown for the representative samples tested at 0.3% and 0.7%. As evident from
Fig. 15(b), micro voids can be seen at crack initiation sites for both specimens. Such micro voids can cause early and multiple
initiations of fatigue cracks, thereby reducing the fatigue resistance of the alloy. Secondary phase particles can also trigger crack
initiation. Striation-like marks, which are indications of crack advancement, can be seen in the slow crack propagation area for the
specimen tested at 0.3%. The crack growth in region II shows features of relatively stable crack growth revealed by the presence of
alternate facets of dark and sharp reflections. Sakara et al. [32] related the bright and dark strips to the mixed mode of cleavage and
dimple fracture. Secondary cracks, which may have been exacerbated by material defect or second phase particles, are visible on both
fracture surfaces. Under 0.7% strain amplitude (Fig. 15(c)), a clear interface between the fatigue crack growth area and fracture
region is obvious. Within the crack propagation area, the so-called tire tracks are present, which is a result of the opposing fatigue
surfaces rubbing each other during propagation. The tire tracks appear to have been created on fatigue debris welded to the specimen
surface due to the rubbing action. The final fracture shows a ductile mode of fracture with multiple dimples.

4. Summary and conclusion

• Microstructural analysis of AISI 410 stainless steel revealed that the alloy exhibits a tempered martensitic structure that is similar
in size and distribution in all the orientations.
• The average size for the grains is found to be around 5 μm.
• Under quasi-static loads, the alloy exhibits the characteristics of a tough alloy with an average tensile strength of 591 MPa, yield
strength of 489 and fracture strain of 21%.
• The alloy displayed a substantial amount of necking, yet possessed low strain-hardening exponent.
• Under strain-controlled fatigue, at all strain amplitudes, cyclic softening was practically observed for the alloy after the first few
cycles until failure.
• The softening degree characterized by softening ratio (SR) is independent of strain amplitude after 0.5% strain amplitude.
• Initial mild cyclic hardening in early fatigue life decreased with increase in applied strain amplitude.
• The shape of the hysteresis loops is generally symmetrical which can be attributed to slip deformation mode.
• Hysteresis loop evolution shows that peak stresses decrease, while plastic strain increases with the number of cycles, an indication
of cyclic softening behavior and accumulating fatigue damage process.
• The alloy is susceptible to dynamic strain ageing effect that is sensitive to strain amplitudes and load direction.
• Essential strain-controlled axial fatigue properties, determined at midlife, resulted in good life prediction, using both Coffin-
Manson and plastic strain energy density equations.
• The fracture surfaces observed under SEM reveal an alloy that is prone to multiple crack initiation both under monotonic and
cyclic loadings. Thus, several slips occurred simultaneously.
• Cracks generally initiated from internal microvoids at the specimens' edges.
• Striation marks are present under low-strain fatigue loading in the crack propagation region while tire tracks are visible under
high strain amplitudes. Thus, the mechanism of crack propagation is influenced by strain amplitude.
• Final fracture is characterized by ductile fracture mode.

10
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

(caption on next page)


11
M.J. Adinoyi, et al. Engineering Failure Analysis 106 (2019) 104166

Fig. 15. SEM micrograph of fractured surfaces of AISI 410 specimens under cyclic axial load.

Declaration of Competing Interest

None.

Acknowledgement

This research is supported by a project grant KACST 230-34, from King Abdulaziz City for Science and Technology, Riyadh, Saudi
Arabia. The authors would also like to acknowledge the support of King Fahd University of Petroleum & Minerals (KFUPM).

References

[1] P. Noury, K. Eriksson, Failure analysis of martensitic stainless steel bridge roller bearings, Eng. Fail. Anal. 79 (2017) 1017–1030, https://doi.org/10.1016/j.
engfailanal.2017.06.035.
[2] M.S. Kumar, M. Sujata, M.A. Venkataswamy, S.K. Bhaumik, Failure analysis of a stainless steel pipeline, 15 (2008) 497–504, https://doi.org/10.1016/j.
engfailanal.2007.05.002.
[3] Y. Li, N. Xu, X. Wu, J. Shi, L. Zhang, M. Zhao, et al., Failure analysis of the 304 stainless steel tube in a gas analyzer, Eng. Fail. Anal. 20 (2012) 35–42, https://doi.
org/10.1016/j.engfailanal.2011.10.004.
[4] H.M. Shalaby, Failure investigation of 321 stainless steel pipe to fl ange weld joint, Eng. Fail. Anal. 80 (2017) 290–298.
[5] S. Kaewkumsai, S. Auampan, K. Wongpinkaew, E. Viyanit, Root cause analysis for 316L stainless steel tube leakages, Eng. Fail. Anal. 37 (2014) 53–63.
[6] G.V. Prasad Reddy, P.M. Dinesh, R. Sandhya, K. Laha, T. Jayakumar, Behavior of 321 stainless steel under engineering stress and strain controlled fatigue, Int. J.
Fatigue 92 (2016) 272–280, https://doi.org/10.1016/j.ijfatigue.2016.07.009.
[7] R. Hormozi, F. Biglari, K. Nikbin, Experimental study of type 316 stainless steel failure under LCF / TMF loading conditions, Int. J. Fatigue 75 (2015) 153–169,
https://doi.org/10.1016/j.ijfatigue.2015.02.014.
[8] X. Yuan, W. Yu, S. Fu, D. Yu, X. Chen, Effect of mean stress and ratcheting strain on the low cycle fatigue behavior of a wrought 316LN stainless steel, Mater. Sci.
Eng. A 677 (2016) 193–202, https://doi.org/10.1016/j.msea.2016.09.053.
[9] W. Chen, X. Fei, Y. Tian, D. Yu, W. Yu, X. Chen, Effect of thermal aging on the low cycle fatigue behavior of Z3CN20.09M cast duplex stainless steel, Mater. Sci.
Eng. A 646 (2015) 263–271.
[10] M. Kamaya, Fatigue properties of 316 stainless steel and its failure due to internal cracks in low-cycle and extremely low-cycle fatigue regimes, Int. J. Fatigue 32
(2010) 1081–1089, https://doi.org/10.1016/j.ijfatigue.2009.12.003.
[11] G. Golanski, S. Mrozinski, Low cycle fatigue and cyclic softening behaviour of martensitic cast steel, Eng. Fail. Anal. 35 (2013) 692–702.
[12] M.C. Tsai, C.S. Chiou, J.S. Du, J.R. Yang, Phase transformations in AISI 410 stainless steel, 332 (2002).
[13] F. Hejripour, D.K. Aidun, Consumable selection for arc welding between stainless steel 410 and Inconel 718, J. Mater. Process. Technol. 245 (2017) 287–299.
[14] B. Vamsi Krishna, A. Bandyopadhyay, Surface modification of AISI 410 stainless steel using laser engineered net shaping (LENSTM), Mater. Des. 30 (2009)
1490–1496, https://doi.org/10.1016/j.matdes.2008.08.003.
[15] C.A. Poblano-salas, J.D.O. Barceinas-sanchez, J.C. Sanchez-jimenez, Failure analysis of an AISI 410 stainless steel airfoil in a steam turbine, 18 (2011) 68–74,
https://doi.org/10.1016/j.engfailanal.2010.08.006.
[16] D.C. Moreira, H.C. Furtado, J.S. Buarque, B.R. Cardoso, B. Merlin, Failure analysis of AISI 410 stainless-steel piston rod in spillway floodgate, Eng. Fail. Anal. 97
(2019) 506–517.
[17] Standard Practice for Preparation of Metallographic Specimens, ASTM E3–01, ASTM International, West Conshohocken, PA, 2001.
[18] Standard Test Methods for Tension Testing of Metallic Materials, ASTM E8/E8M-08, ASTM International West, Conshohocken, PA, 2008.
[19] Standard Practice for Strain-Controlled Fatigue Testing, E 606–99, 24 ASTM International, West Conshohocken, PA, 2004 (1999).
[20] W.D. Callister, D.G. Rethwisch, Materials Science and Engineering, 8th ed., John Wiley and Sons, Inc, 2011.
[21] R.I. Stephens, A. Fatemi, R.R. Stephens, H.O. Fuchs, Metal Fatigue in Engineering, 2nd ed., 3 John Wiley & Sons, 2001.
[22] R. Branco, J.D. Costa, F.V. Antunes, Low-cycle fatigue behaviour of 34CrNiMo6 high strength steel, Theor. Appl. Fract. Mech. 58 (2012) 28–34, https://doi.org/
10.1016/j.tafmec.2012.02.004.
[23] V. Mazánová, V. Škorík, T. Kruml, J. Polák, Cyclic response and early damage evolution in multiaxial cyclic loading of 316L austenitic steel, Int. J. Fatigue 100
(2017) 466–476, https://doi.org/10.1016/j.ijfatigue.2016.11.018.
[24] M. Hornqvist, B. Karlsson, Dynamic strain ageing and dynamic precipitation in AA7030 during cyclic deformation, Procedia Eng. 2 (2010) 265–273.
[25] C. Chen, B. Lv, F. Wang, F. Zhang, Low-cycle fatigue behaviors of pre-hardening Hadfield steel, Mater. Sci. Eng. A 695 (2017) 144–153.
[26] Z. Huang, D. Wagner, C. Bathias, Some metallurgical aspects of dynamic strain aging effect on the low cycle fatigue behavior of C – Mn steels, Int. J. Fatigue 80
(2015) 113–120, https://doi.org/10.1016/j.ijfatigue.2015.04.008.
[27] Z.Y. Huang, D. Wagner, Q.Y. Wang, M.K. Khan, J.L. Chaboche, A low cycle fatigue model for low carbon manganese steel including the effect of dynamic strain
aging, Mater. Sci. Eng. A 654 (2016) 77–84, https://doi.org/10.1016/j.msea.2015.12.022.
[28] A. Sarkar, A. Nagesha, P. Parameswaran, R. Sandhya, M. Okazaki, Implications of dynamic strain aging under LCF-HCF interactions in a type 316LN stainless
steel, Mater. Sci. Eng. A 708 (2017) 91–103, https://doi.org/10.1016/j.msea.2017.09.057.
[29] P. Verma, G.S. Rao, P. Chellapandi, G.S. Mahobia, K. Chattopadhyay, N.C.S. Srinivas, et al., Dynamic strain ageing, deformation, and fracture behavior of
modified 9Cr – 1Mo steel, Mater. Sci. Eng. A 621 (2015) 39–51, https://doi.org/10.1016/j.msea.2014.10.011.
[30] N.C.S. Srinivas, P. Verma, V. Singh, Dynamic strain ageing behaviour of modified 9Cr-1Mo steel under monotonic and cyclic loading, Procedia Eng. 184 (2017)
765–772, https://doi.org/10.1016/j.proeng.2017.04.155.
[31] V. Shankar, A. Kumar, K. Mariappan, R. Sandhya, K. Laha, A.K. Bhaduri, et al., Occurrence of dynamic strain aging in alloy 617M under low cycle fatigue
loading, Int. J. Fatigue 100 (2017) 12–20, https://doi.org/10.1016/j.ijfatigue.2017.03.001.
[32] S. Sarkar, C. Siva Kumar, A. Kumar Nath, Effect of mean stresses on mode of failures and fatigue life of selective laser melted stainless steel, Mater. Sci. Eng. A
700 (2017) 92–106, https://doi.org/10.1016/j.msea.2017.05.118.

12

You might also like