You are on page 1of 36

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/313593748

Characteristics and damage mechanisms of bending fretting fatigue of


materials

Article  in  International Journal of Damage Mechanics · February 2017


DOI: 10.1177/1056789517693412

CITATIONS READS

12 1,785

4 authors, including:

Mohammad Asaduzzaman Chowdhury Arefin Kowser


Dhaka University and Technology, Gazipur Dhaka University of Engineering & Technology
111 PUBLICATIONS   1,281 CITATIONS    52 PUBLICATIONS   195 CITATIONS   

SEE PROFILE SEE PROFILE

Quazi Md. Zobaer Shah


Healthcare Pharmaceuticals Ltd
10 PUBLICATIONS   38 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Structural Health Monitoring of Mechanical Structures View project

Improvement of Interfacial Adhesion Performance of the Kevlar Fiber Mat by depositing SiC/TiO2/Al2O3/ Graphene Nanoparticles View project

All content following this page was uploaded by Arefin Kowser on 06 March 2018.

The user has requested enhancement of the downloaded file.


Review
International Journal of Damage
Mechanics
0(0) 1–35
Characteristics and damage ! The Author(s) 2017
Reprints and permissions:
mechanisms of bending fretting sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1056789517693412
fatigue of materials journals.sagepub.com/home/ijd

Mohammad Asaduzzaman Chowdhury,


Md. Arefin Kowser, Quazi Md. Zobaer Shah and
Suman Das

Abstract
Fretting fatigue has attracted substantial research interest in recent decades owing to its relevance in a
wide range of applications. This paper reviews previous studies and describes ongoing research.
Particularly, only a few studies on bending fretting fatigue have been conducted. This review paper
emphasizes the effect of bending fretting fatigue of different materials under different operating param-
eters. In addition, the damage mechanisms with respect to nature of failure of materials are also discussed.
This paper can be used as a reference for design and development of modern technologies and selection
of appropriate material in industries.

Keywords
Bending fretting fatigue, materials, operating conditions

Introduction
Fretting, which is gradual wear of something by rubbing or gnawing, is a wear process that
occurs at the contact area between two materials under load and subject to minute relative
motion by vibration or some other forces (Hattori et al., 1988). Although several articles on
fretting fatigue are available in the literature, research on bending fretting fatigue is limited.
Fretting fatigue occurs when one body in contact with another tries to slide or roll over that
body. Hattori et al. (1988) and King and Lindley (1981) proved that fatigue strength during
fretting decreases to less than one-third of that without fretting. Bending fretting fatigue occurs

Dhaka University of Engineering and Technology, Gazipur, Bangladesh


Corresponding author:
Mohammad Asaduzzaman Chowdhury, Dhaka University of Engineering and Technology, Gazipur 1700, Bangladesh.
Email: asadzmn2014@yahoo.com
2 International Journal of Damage Mechanics 0(0)

when the relative movement is influenced by cyclic bending load. Common examples are found
in contact pairs, bolts, railway axles, turbine dove-tail-like blades, nuclear reactors, pressure
vessels, and overhead electrical conductors (Ebara and Fujimura, 2006; Hills and Nowell, 1994;
Neu et al., 2000; Mayama et al., 2008; Nowell et al., 2006). The initiation and propagation of
fatigue cracks are severely characterized by the microstructure of materials. The dislocation
seriously affects the plastic deformation, and the relationship between dislocated structures and
plastic deformation is a fundamental subject of concern (Yaguchi et al., 2001). A small number
of studies (Gaudin and Feaugas, 2004; Kubota et al., 2011; Yaguchi et al., 2001) have demon-
strated the relationship between plain fatigue and dislocated structure of materials. The study
in Peng et al. (2011) and Gosar and Nagode (2015) focused on the role of ecological hydrogen
on the fretting fatigue quality of austenitic stainless steel. The relationships between the reach-
ing surfaces and little cracks within the materials have been investigated, including the role of
hydrogen gas. A 316L austenitic stainless steel was taken as the sample because it possesses the
crystal structure of face centered cubic (FCC) lattice, and its dislocation configurations are easy
to locate (Zalnezhad et al., 2013; Zhu et al., 2016). An experimental evaluation was conducted
to investigate the fretting fatigue life of multilayer Cr–CrN-covered AL7075-T6 alloy samples
with high adhesion strength of the substrate. In fact, the bond quality of coating stands out
among the most basic issues in magnetron sputtering system (Luo, 2013; Peng et al., 2013).
Previous research (Peng et al., 2014) investigated the bending fretting fatigue of 7075 aluminum
alloy. The work focused on behaviors of bending fretting fatigue, which are affected by the
applied bending stress and the number of cycles on the same ratio. The objective was to
identify the life reduction caused by the fretting induced by the pure bending load, and to
determine the relationship between the bending fretting fatigue behavior and dislocation struc-
tures of microstructures. The behavior of soft bending fretting fatigue of medium carbon steel
(LZ50 steel), aluminum alloy (7075) (Peng et al., 2014), and stainless steel (316L steel)
(Zalnezhad et al., 2013) have been investigated (Majzoobi and Minaii, 2013). Recently,
30CrNiMo8 and 17CrNiMo6 alloy steels were investigated under the same conditions. The
impact of deep rolling on the bending fretting fatigue resistance of Al7075 is studied using
the rotating bending fretting fatigue tests conducted on a Moore rotary bending mechanical
assembly (Zalnezhad et al., 2014). A test assessment by a rotating bending mechanical assembly
(Song et al., 2014) was conducted to investigate the fretting fatigue life of multilayer Cr–CrN-
covered AL7075-T6 composite samples with higher adhesion quality to substrate, given that the
covering grip quality is vital among the most basic issues in magnetron sputtering system (Peng
et al., 2014). Fretting damage failure of a Chinese carbon railroad axle RD2 was investigated
by using a small-scale axle-test apparatus and recreating tests at different rotator rates of 1800
and 2100 r/min (Gutkin and Alfredsson, 2008). Rotating bending was studied for crack growth
in a press-fitted test rig (Savaria et al., 2016; Yang et al., 2014; Yuan et al., 2014). In recent
years, a number of studies on bending fatigue (Dao and Vu, 2016; Goidescu et al., 2015;
Haidyrah et al., 2016; Hwang et al., 2016; Kamaya, 2016; Li et al., 2016; Ma et al., 2015)
and fretting fatigue have been reported (Foletti et al., 2016; Huang et al., 2016a, 2016b; Shi
et al., 2016; Winkler et al., 2015), whereas comparatively less studies were performed on
bending fretting fatigue (Ding et al., 2014; Johnson, 1995; Liu et al., 2013). In the present
study paper, test rigs for bending fretting fatigue are classified into reciprocating and rotary
types, which are subdivided into branches. The experimental setup and testing were discussed,
and cracks, scars and fractured surfaces were evaluated using optical microscopy, transmission
electron microscopy, scanning electron microscopy.
Chowdhury et al. 3

Figure 1. Fretting fatigue test method; shapes and sizes of the sample and contact pad (dimensions are in mm):
(a) fatigue test sample and (b) bridge type contact pad (Kubota et al., 2011; Peng et al., 2011).

Reciprocating bending fretting fatigue test rigs


Bar-spring-tightened double-bending fretting fatigue test rig (Kubota et al., 2011; Peng et al.,
2011)
Figure 1 shows the fretting fatigue test strategy utilized as a part of the test. Two contact pads were
squeezed onto the front and rear surfaces of the fatigue test sample. The contact load was adopted to
maintain the nominal pressure at 100 MPa. The contact load was applied by fixing the bar springs
with cinching bolts. The unbending nature of the bar springs was thoroughly outlined so that the
avoidance when the contact load achieved the predefined value was adequately extensive compared
with the string lead of the clasping bolt. This setup ensured the close conformity of the contact load.
The fine string of the clipping bolt also added to the exact conformity of the contact load. Although
4 International Journal of Damage Mechanics 0(0)

contact load may decrease caused by the improvement of fretting wear, the contact load was steady
during the test, as the fretting wear was negligible compared with the change in the bar springs.
After the fretting fatigue test, the actual decrease in contact load was under 5% of the starting load.
The fatigue loading was in-plane bending. The fretting fatigue test was performed with stress ratio of
1 at frequency of 18.7 Hz at ambient temperature. The maximum nominal pressure of the sample
was measured by a strain gauge glued to the sample surface. The shapes and sizes of the fatigue test
sample and the contact pad are shown in Figure 1. The contact surfaces of the sample and the
contact pad were buffed to provide similar fretting damage and little fretting fatigue cracks.
A square bar sample was utilized to obtain high level of surface contact. An extension contact
pad was utilized to gauge the tangential power. As shown in Figure 1, a strain gauge was glued
at the focal point of the recessed part of the contact pad. The relationship between the yield of the
strain gauge and tangential power was determined by limited component examination. The tangen-
tial power coefficient was characterized as the proportion of the tangential power to the contact
power. The test conditions were 0.12 MPa hydrogen gas and ambient air. The purity of the hydrogen
gas was greater than 99.9999%. Before the fretting fatigue test in the hydrogen gas, the gas chamber
was cleared and then infused with nitrogen gas for three times to increase the purity of hydrogen gas.
The accomplished vacuum was greater than 5  103 Pa.

Shoe and bolt-tightened bending fretting fatigue test rig (Ebara and Fujimura, 2006)
Fretting fatigue tests for Ti–6Al–4V compound were conducted by using a fretting fatigue
sample with a bolt-tightened shoe on both sides of the plate. The 10-mm-thick fretting fatigue
sample plate is shown in Figure 2. Samples were cut from a round pole 240 mm long and 60 mm
in wide. Figure 3 shows the fretting shoe and bolt. The fretting shoe and the bolt were produced
using Ti–6Al–4V compound and SCM4 steel, respectively. Contact pressure was applied point S
in Figure 2. Contact load was obtained from the strain measured by a strain gauge located at
the bolt center. The contact between fretting fatigue sample and shoe is similar to that between a
plate and a disc. The contact width is too narrow compared with the plate width. Consequently,
the most extreme contact stress was calculated using the Hertz formula. The maximum load
Pmax is 0.418(PE/lR)1/2, where P is the contact load per unit width (MPa), E is Young’s modulus

Figure 2. Fretting fatigue sample (Ebara and Fujimara, 2006).


Chowdhury et al. 5

Figure 3. Fretting shoe and bolt (Ebara and Fujimara, 2006).

Figure 4. Fretting fatigue testing view (Ebara and Fujimara, 2006).

of the shoe (MPa), R is radius of the shoe, and l is the contact length. The plane-bending stress
was the corrected stress on the S spot determined by the strain picked up from a strain gauge
located on the G point. A plane bending fatigue testing machine (4.9 kN) was utilized. The
testing recurrence was 13.3 Hz. Figure 4 shows the testing view of fretting stress. The stress
ratio was taken 0.05 in their experiment.

Servo hydraulic clamped bending fretting fatigue test rig (Peng et al., 2014; Majzoobi and
Minaii, 2013)
A commercial servo hydraulic uni-axial fatigue tester (SHIMADZU EHF-UM100K2 servo fatigue
machine) was utilized for the bending fretting fatigue tests. The test rig shown in Figure 5 was
mounted on the platform of the fatigue machine. An attaching bolt was utilized to force the ordinary
load to the plate sample at A–B line and a point (Figure 6) during the tests. On the water-driven
servo fatigue test machine, through the new arrangement, two contact setups (i.e., line and point
contacts) of bending fretting fatigue can be achieved. The ordinary load was measured and
6 International Journal of Damage Mechanics 0(0)

Figure 5. Two contact configurations of bending fretting (Peng et al., 2013; 2014).

Figure 6. Bending fretting fatigue specimen geometry. (a) 30CrNiMo8, (b) 7075 alloy, (c) LZ50 steel, and (d) 316L
steel fatigue test rig. (a) and (b) point contact and (c) and (d) line contact (Peng et al., 2013; 2014).

controlled through a little load cell, as shown in Figure 5. The fatigue tests changed to the plain
bending fatigue when no ordinary load was forced. The bending load was connected at the site of C–
D line, as shown in Figure 6 for all tests. The bending load connected to the sample brought on a
change in the center, the sample was bent downwards, resulting in a tangential power that is
transmitted between the pad and the sample. The patterned bending loads (W) were forced at the
C–D line and B point in Figure 6 by the servo fatigue machine. For the bending fatigue tests,
maximum repeating bending loads were calculated using the following equations

a, max ¼ 6WL=bh2


a, max ¼ 32WL=d3

where  a,max is the maximum bending stress, W is the cyclical bending load, L is the distance between
bending load and the fretting pads, b is the width of the sample, h is the height of the sample, and d is
the diameter of the sample.
Normal stresses and loads were imposed by fretting pads and calculated based on Hertzian
contact theory (Johnson and Greenwood, 1997; Lee and Mall, 2006)]. Stress ratio R and frequency
(in Hz) were determined for definite frequency and sinusoidal loading.
Chowdhury et al. 7

Figure 7. Schematic of fretting fatigue test rig: (a) fretting ring and (b) fretting fatigue test machine (Peng et al., 2013;
Zalnezhad et al., 2013).

Rotary bending fretting fatigue test rig


Light-/general-purpose RoBFFTR (Peng et al., 2013; Zalnezhad et al., 2013). Figure 7(a) shows
a schematic of the fretting fatigue test rig, which was built in-house for this study. A rotating
bending fretting fatigue test machine was connected at a frequency of 50 Hz and steady contact
pressure of 100 MPa at room temperature, as shown in Figure 7(b). This sort of testing was picked in
light of the fact that it delivers the best measure of stress on the sample surface (Peng et al., 2013).

Thick disc type pads/press-fitted joint type RoBFFTR (Gutkin et al., 2008; Savaria et al.,
2016). The device utilized for the analyses is shown in Figure 8 (left). The assembly constituted
a pole/shaft shrink-fitted into a packaging, that is, the pole/shaft was cooled and the packaging
heated to encourage the assembly process. The free compelling of the pole/shaft was embedded into
an orientation and a bending weight of 20 kg was connected, see Figures 8. The electrical engine that
was utilized to rotate the assembly was joined with the more slender part of the packaging in Figure
8 by a binding method.

Medium purpose RoBFFTR (Moore fatigue testing machine) (Majzoobi and Minaii, 2013;
Zalnezhad et al., 2014). The bending fatigue procedure can be performed by the Moore fatigue
testing machine. The upside of the Moore gadget is that the samples are free of shear stresses. As the
sample turns, any point on the sample surface continuously encounters tensile and compressive
stresses in every cycle. With a specific end goal to apply and measure the contact stress, a proving
ring was designed. The ring assembly is shown in Figure 9. The screw was tightened to compress the
spring; also, the power in the spring is determined. Fretting fatigue tests were directed on a Moore
rotating bending machine at 30 Hz frequency. The contact power was 1300 N and stayed consistent
for all tests. The assembly of the sample, pads, and proving ring, mounted on a Moore machine, is
shown in Figure 10.

Complex/specific (railway axle) purpose RoBFFTR (Gutkin and Alfredsson, 2008; Song et al.
2014). A rotary bending fretting fatigue test framework was developed, as shown in Figure 11. A
press-fitted wheel-axle structure (15) assembly was utilized for this test rig. The axle sample was
8 International Journal of Damage Mechanics 0(0)

Figure 8. (Left) Shrink-fit sample used in the experiments in Mdzimba (2006), and Gosar and Nagode (2015).
Schematic of the experiments including load and movements (right) (Gutkin et al., 2008; Savaria et al., 2016).

Figure 9. Ring load cell (a) position of the strain gauge and (b) position of the calibrating spring (Majzoobi and Minaii
2013; Zalnezhad et al., 2014).

mounted on the rig by a pinchcock (16) to obtain the rotary movement. The rotating speed range
was 0 to 5000 r/min, and the rotating velocity of the engine was controlled by an AC converter with
control precision of 1 r/min. A driven wheel (20) was gathered on the base to support the press-fitted
wheel-axle and pivot with it during the experimental process. The area of the bolster wheel can be
modified as per the experimental process. During the test process, the driven wheel was affixed on
the base using eight bolts. A high-speed bearing was utilized to supply a bending load on the half
wheel seat sample. The distance from the bearing focus to the press-fitted contact edge was similar to
Chowdhury et al. 9

Figure 10. Arrangement of sample, pads, proving ring, and Moore machine (Majzoobi and Minaii 2013; Zalnezhad
et al., 2014).

Figure 11. Rotary bending fretting fatigue test rig. 1: motor; 2: bracket for motor; 3: surface plate; 4: base; 5: oil
box; 6: cooling oil; 7: oil pump; 8: oil pipe; 9: tachometer; 10: revolution counter; 11: bracket for sensor; 12: bracket
for axle; 13: loading force; 14: axle sample; 15: half wheel set sample; 16: pinchcock; 17: headstock; 18: principal axle;
19: coupling; 20: driven wheel; 21: driven shaft; 22: oil pipe; 23: pressure sensor; 24: control chamber; and 25: cable
(Gutkin and Alfredsson, 2008; Song et al., 2014).
10 International Journal of Damage Mechanics 0(0)

the distance on the real axle. The bending load was generated by a dead weight through a lever
framework which was not shown in Figure 11.

Evaluation of S–N curves


S–N curves for different materials under different test rigs show specific characteristics because of
material properties and fretting actions. Fretting fatigue lifetime was unmistakably shorter than that
of plain fatigue when the same cyclic bending stresses were applied.
For servo-hydraulic fretting fatigue (Peng et al., 2013; 2014), the fatigue lifetime demonstrates a
seriously diminishing prominent pattern that is much shorter than that of plain bending stress under
the same repeating load. As the bending stress expands, the distinction of the lifetime between plain
fatigue and fretting fatigue gradually diminishes because of the bending stress expanding close to the
yield stress of the material. In the event of the same load magnitude, fretting fatigue lifetime ser-
iously diminishes with the addition of bending stress up to a critical point, and then the fretting
fatigue lifetime demonstrates a negative behavior to the bending stress increase taking a state of
‘‘C’’-type S–N curves. Zhou and Zhu (2004) indicated illustrative exploration result. As the bending
stress approaches the value of the yield stress (for instance, Figure 12(a), (b) and (d)), the fretting
fatigue lifetime diminishes significantly once more. Considering its relatively weaker yield quality,
the decrease for the second time cannot be observed for medium carbon steel (LZ50 steel). By after-
effects of various materials, the S–N curve shows a shape close to the temperature–time transform-
ation curve of composite steels in the isothermal cooling process, similar to the Greek letter ‘‘"’’ as
shown in Figure 13. From Figure 13, the base lifetime of fatigue is outlined and crack starts effort-
lessly and propagation occurs in the mixed fretting regime (MFR), similar to the lower nose of the ‘‘"
curve.’’ On the off chance that bending stresses decrease, the long fatigue lifetime and a little wear
develop in the partial slip regime (PSR). For genuine mechanical applications, the bending stress
should be managed at a lower level in the PSR to increase wealthier administration life and depend-
ability. For a sufficient safety coefficient, the well-being space must be below the S–N curve of the
PSR (see Figure 13). In the slip regime (SR), however, the fatigue lifetime is higher than that in the
MFR; it is not safe on account of the high bending stress near the yield quality of materials. With the
increment of typical load, the state of ‘‘"-curve’’ stays steady yet the area of the nose was changed
clearly as shown in Figure 14. The most limited lifetime develops under the medium typical load
(500 N). Figure 15 demonstrates that the lifetime of bending fretting fatigue of 17CrNiMo6 steel
develops as capacity of the ordinary load.
In the case of Ti–6Al–4V alloy (shoe and bolt type rig) (Ebara et al., 2006; Nowell et al., 2006),
shown in Figure 16, the relation between plane-bending stress and stress over the contact area where
fretting fatigue crack begins is equivalent to 1284 MPa. The bigger the contact stress, the more
effortless the fretting fatigue crack starts at the same bending stress. The connection between plane-
bending stress and stress over the contact area can be arranged into three zones, for example, region
A where fatigue failure occurs, territory B where fretting fatigue failure occurs and zone C where no
fatigue and fretting fatigue failure occur over 107 cycles. The plane-bending stress at the contact
region where fretting fatigue failure starts was found to be directly diminished as stress over the
contact area expanded.
For SUS 304 in H2 environment (bar spring tightened test rig) (Kubota et al., 2011; Peng et al.,
2011), the S–N graph of is shown in Figure 13. Although a small number of information focuses was
adopted, the measurements were adequate to affirm the lessening in the fretting fatigue strength
caused by hydrogen gas and to determine the stress amplitude utilized in the two-stage test. In the
limited life region, where the stress amplitude was higher than that of the fretting fatigue limit in air,
Chowdhury et al. 11

Figure 12. S–N curves of the bending fretting fatigue for various materials: (a) 30CrNiMo8 alloy steel, (b) 7075
aluminum alloy, (c) LZ50 steel, and (d) 316L austenitic stainless steel (Peng et al., 2014; Zalnezhad et al., 2014).

Figure 13. S–N curve of bending fretting fatigue: PSR, MFR, and SR (Peng et al., 2013; 2014).
12 International Journal of Damage Mechanics 0(0)

Figure 14. S–N curve of bending fretting fatigue of 7075A alloy under carried normal loads (Peng et al., 2013; 2014).

Figure 15. Fatigue life time of bending fretting fatigue of 17 CrNiMo6 steel as a function of the normal load (Peng et
al., 2014; Majzoobi and Minaii, 2013).

600
Stress
400
Stress,S

200

0
4.2
1000
320
25
770
1000
59
1000
180
1000

N,Cycle

Figure 16. S–N curve (left), and relation between plane-bending stress and stress over the contact area (right)
(Ebara et al., 2006; Nowell et al., 2006).
Chowdhury et al. 13

Figure 17. Effect of hydrogen gas on fretting fatigue limit in hydrogen gas in prestrained SUS304 (Kubota et al.,
2011; Peng et al., 2011).

Figure 18. (a) S–N curve of plain and fretting fatigue for uncoated samples and (b) S–N curve for uncoated and
Cr–CrN-coated with highest surface adhesion samples (Song et al., 2014; Zalnezhad et al., 2013).

the fretting fatigue life was longer in the hydrogen gas than in air. The nucleation of little fretting
fatigue cracks lesser than 10 mm was within a couple percent of the fretting fatigue life for both the
hydrogen gas and air situations. This procedure of nucleation of microcracks and their capture was
repeated until a decrease of almost 90% of the fretting fatigue life was achieved. In this way, the
motivation behind why the fretting fatigue life in hydrogen gas was longer than that in air was the
postponement of the start of the steady crack propagation.
For Cr–CrN coated AL7075-T6 (simple rotary bending test rig) (Majzoobi and Minaii, 2013;
Zalnezhad et al., 2013), Figure 18(a) demonstrates the quantity of cycles to failure versus bending
stress for plain fatigue and fretting fatigue of uncoated samples. The fretting fatigue crack formed in
the area where the frictional shear stress on the contact surface privately focused. The fatigue life of
uncoated samples was reduced with the increase in bending stress. The fretting deleteriously affects
fatigue life. The S–N curve of fretting fatigue for uncoated and Cr–CrN-covered samples with the
most astounding surface attachment is delineated in Figure 18(b). The fretting fatigue lives of Cr–
CrN-covered samples were improved under both low and high cyclic fatigue compared with that of
uncoated samples.
14 International Journal of Damage Mechanics 0(0)

Figure 19. (a) S–N curves of rotary bending fretting fatigue for the small-scale axle samples and (b) fretting fatigue
life gap between the simulation speed of 1800 and 2100 r/min (Majzoobi and Minaii, 2013).

For Chinese carbon railway axle RD2 (complex type test rig) (Gutkin and Alfredsson, 2008; Song
et al., 2014), the fatigue strength of the samples was largely diminished by the impact of fretting. As
shown in Figure 19(a), the S–N curves indicate a nonlinear conduct. As a rule, the fatigue life of tests
under 2100 r/min was shorter than that of 1800 r/min. However, the crevice between them differs
with the applied load level, as seen in Figure 19(b). At the low bending fatigue stress level under
120 MPa, samples under both test conditions come up short on the fatigue limit (107 cycles). In the
medium bending stress level, the life of each sample under 2100 r/min was shorter than that under
1800 r/min. Whenever the applied bending fatigue load was sufficiently high (>200 MPa), the fatigue
lives for both two running velocities (1800 and 2100 r/min) were close (verging on equivalent), as
shown in Figure 19(a) and (b).
In the case of fatigue resistance of Al7075 (Moore test rig) (Majzoobi and Minaii, 2013; Zalnezhad
et al., 2014), various plain and fretting fatigue tests were conducted on the non-rolled samples.
Different fretting conditions radically decrease the fatigue life of the samples, as shown in Figure
20. The fatigue tests were conducted with a frequency of 30 Hz and under a consistent typical
contact force of 1300 N. Under 190 MPa, the fatigue life decreased from 1,300,000 cycles for
plain fatigue to approximately 65,000 cycles for fretting fatigue conditions. This means a decrease
of approximately 95% in plain fatigue life compared with FFL.

Evaluation of fracture and fretting damage zones (Peng et al., 2014; Majzoobi
and Minaii, 2013)
Figure 21 shows the common morphologies of the contact surface of bending fretting fatigue. For
the PSR region, an asymmetrical wear annulus slightly molded in the damage zone is observed.
Slight oxidation as well as some furrowing can be found. At the side of contact near stacking end
side, some oxidative debris heaped up. Contact interfaces were facilitated by the flexible disfigure-
ment where no small scale cracks were found. The wear system in the PSR for bending fretting
fatigue is a mix of abrasive wear and slight oxidative wear. In Figure 22, the advancement of the
wear scratches are picked up by various samples after various cycles. At the early stage
(N ¼ 104cycles) of the test under a stress of 305.6 MPa, an annular morphology was shaped because
of coupling of the incomplete slip and gross slip between the contact interface.
Chowdhury et al. 15

Figure 20. S–N curves for plain and fretting fatigue conditions (Majzoobi and Minaii, 2013; Zalnezhad et al., 2014).

Figure 21. Morphologies of fretting damage zones of bending fretting fatigue for 7075 alloy: PSR- a,max ¼ 241.9 MPa,
N ¼ 106; PSR- a,max ¼ 241.9 MPa, N ¼ 106; SR- a,max ¼ 814.9 MPa, N ¼ 2  105 (Peng et al., 2014; Majzoobi and Minaii,
2013).

Abrasive wear is the primary culprit for this early damage stage. As the quantity of cycles
expands, the damage accumulated with the augmentation of fretting scratch size, profound furrow-
ing grooves, slight separation, and some piled-up debris. No crack can be captured in the fretting
scar, even at N ¼ 5  104 cycles. To further build the no of cycles to 1  105, the furrowing grooves
persistently developed, and delaminating separation was turned into the prime concern at the
stacking end side with some oxidative debris particles. It demonstrates that an obvious crack was
recognized by SEM at N ¼ 1.5  105 cycles. The sample achieves its bending fretting fatigue life,
which is set apart by the propagation of a full scale crack. It demonstrates that a more serious
separation range developed as a component of the cycles can be observed. At the same time, the
delamination territory clearly expanded and compared with a more extreme separation. In this
manner, the damage systems of fretting contact zone in the MFR are grating wear, oxidative
wear, and delamination, and cracking. A physical model shown in Figure 23 can graph the
16 International Journal of Damage Mechanics 0(0)

Figure 22. Evolution of damage surface of contact zones in the MFR as function of the number of cycles (Peng et al.,
2014; Majzoobi and Minaii, 2013).

Figure 23. Physical model of damage mechanisms in the MFR (Peng et al., 2014; Majzoobi and Minaii, 2013).

advancement of damage procedures in the MFR. In the SR, more serious wear occurs; however, no
surface crack can be found on the fretting scars. Like the MFR, the wear systems in the SR do not
change, that is, the instruments are rough wear, oxidative wear, and delamination. As the repeating
bending stress connected with ordinary load is applied, the relative removal between the sample
contact range and upper fretting pad results in the plastic distortion and work hardening in the
contact region. It causes the material to be embrittled and facilitates development of a white layer
(Chuanfu et al., 1992). The contact zone material surface experiences elective tangential powers
Chowdhury et al. 17

Figure 24. (a) SEM image of fretting damage zone: W ¼ 5.75 KN,  a,max ¼ 532.4 MPa, N ¼ 107 cycles (Gosar and
Nagode, 2015). (b) SEM image of the cross-section of the fretting damage zone: W ¼ 6.25 KN,  a,max ¼ 578.7 MPa, and
107 cycles (Peng et al., 2011; Gosar and Nagode, 2015).

influenced by bending fatigue stress, the material is withdrawn to frame the debris in the intense
damage zone. The plastic stream and delamination likewise develop, and then the withdrawn par-
ticles are ground and oxidized over and over again to form the fine oxidative debris bed. Finally, the
debris would be moved to the outside of the contact boundary.

SEM view of SUS-316L (Peng et al., 2011; Gosar and Nagode, 2015)
With the increase in bending fatigue load, the fretting contact zone exhibited wear and tear, as well
as produced minuscule cracks in the network, which was opposite to the course of the small scale
slip, as shown in Figure 24. The cracks might be spread in further cycles and initiate cracking of
the sample. Some furrowing grooves were found as shown in Figure 24. The damage levels of the
contact zones were emphatically indigent on the advancement of the quantity of cycles. At the point
when the greatest cyclical bending load achieved 6.0 KN ( a,max ¼ 555.56 MPa), the sample would
break at the number of cycles under 107. The SEM pictures of the cross-area (Figure 25) show that
the tiny crack started along the path of approximately 30–45 to the surface toward the starting
stage (Figure 25(b)), and propagates opposite to the contact surface with the increase in the quantity
of the cycles. The crack initiation took a sudden move to an opposite course of approximately
90 mm long. Moreover, the crack engendering was chiefly transgranular (Figure 24(b)).
Figure 25(b) demonstrated that the depth of the crack was approximately 72 mm, which was
between the depth of the most extreme Hertz stress (45.37 mm) and the depth of the contact stress
impact territory (approximately 113 mm as determined by Hertz contact hypothesis). The phase of
start and propagation of the angled crack was controlled by the local contact stress of fretting, and
the opposite phase of the propagation was controlled by the bending fatigue stress. The start of the
crack was usually subject to the nearby fretting action and the propagation was subject to the entire
fatigue. At the point when the cyclical bending stress expanded further to a higher level, the lifetime
of the fretting fatigue in this manner was shortened; nevertheless, the element of two stages was still
unaltered (Ding et al., 2014).
18 International Journal of Damage Mechanics 0(0)

Figure 25. SEM images of the cross-section of the fretting damage zone: W ¼ 6.0 KN,  a,max ¼ 555.6 MPa, and 107
cycles (Zalnezhad et al., 2013).

Fracture analyses (Peng et al., 2011; Majzoobi and Minaii, 2013)


Figure 26 shows typical crack surfaces of various materials under fluctuated test parameters. The
bending fretting fatigue cracks obviously start in the subsurface at approximately 100 mm under-
neath the surface, where the contact stress is maximal. The bending fretting cracks more usually start
at the destinations where several impurities or secondary phases exist. From the earliest starting
point of break source, crack propagating toward test focus shows spiral components (i.e., the crack
propagates along the outward direction), which are consistent with the heading of crack propaga-
tion, and are opposite to fatigue striations. With the increase of bending stress, the auxiliary cracks
are shaped by the high fatigue stress, and the size and measure of the auxiliary cracks with an
intermittent dissemination clearly increases. During the crack propagation, the local stress would be
casual because of the optional crack arrangement. In this manner, the fatigue crack engendered on
the strip component (Iwabuchi et al., 1983).
At the point when the bending stress was sufficiently high (W ¼ 6.25 KN,  a,max ¼ 578.7 MPa), the
secondary cracks opposite to the crack surface are observed (Figure 27(a)). With further increase in
bending stress, the size and measure of the secondary cracks expanded clearly (Figure 27(b)). The
numerous free crack sources can be created under higher stress levels. The local stress would be
constant during the propagation of the secondary cracks, connecting to the principal crack. In any
case, this process was not a vital part of the damage.
Figure 28 exhibits the actual mechanism of the crack propagation for 7075 aluminum alloy and
30CrNiMo8 composite steel. The crack propagation can be partitioned into three stages: (i) Stage I
(crack start stage), the course of crack propagation is at a specific point (601) to the contact
surface, which is controlled by the local contact stresses, and the primary crack connections to
the surface grade cracks, which propagate alone. (ii) Stage II (intermediate move stage), as the
depth surpasses the maximal Hertzian contact depth (Foletti et al., 2016; Huang et al., 2016), the
crack propagation is both ruled by the contact stress and plain bending fretting stress, the crack
initiation changes to a higher edge. (iii) Stage III (normal propagation stage), the crack propagation
is controlled only by the plain bending fretting stress, and the propagation direction is opposite to
the contact surface; its propagation behavior agrees with the plain fatigue. The two propagation
stages controlled by the fretting contact stress are the principal variables influencing the fatigue life.
Chowdhury et al. 19

Figure 26. Morphologies of fracture surface and fatigue sources for different materials: (a) 30CrNiMo8 steel,
 a,max ¼ 821 MPa; (b) 7075 alloy,  a,max ¼ 305.6 MPa; (c) LZ50 steel,  a,max ¼ 509.3 MPa, and (d) 316 L steel,
 a,max ¼ 578.7 MPa (Peng et al., 2014; Majzoobi and Minaii, 2013).

Figure 27. SEM images of the fretting fracture surface: (a) W ¼ 6.25 KN,  a,max ¼ 578.7 MPa and (b) W ¼ 6.5 KN,
 a,max ¼ 601.9 MPa (Peng et al., 2011; Zalnezhad et al., 2013).

The entire bending fretting fatigue crack propagation can be described by utilizing a physical model,
as shown in Figure 29.

Microstructure evolution on TEM (Peng et al., 2011; Majzoobi and Minaii, 2013)
In terms crystal structure, 316L austenitic stainless steel and 7075 aluminum combination both have
FCC structures. Nevertheless, the developments of disengagement setup for 316L steel and 7075
20 International Journal of Damage Mechanics 0(0)

Figure 28. Crack propagation paths for different materials: (a) 7075 alloy,  a,max ¼ 331.0 MPa, 1.5  105 and
(b) 30CrNiMo8 steel,  a,max ¼ 746 MPa, N ¼ 1.8  105 (Peng et al., 2014; Majzoobi and Minaii, 2013).

Figure 29. Crack propagation model with three stages (Peng et al., 2014; Majzoobi and Minaii, 2013).

alloy are entirely different because of the difference in stacking flaw vitality. Hence, different crack
start components are developed.
Twinning nucleation model for the FCC-structured metals with low-level stacking fault energy
(e.g., austenitic stainless steel). The disengagement thickness of tempering 316L austenitic stainless
steel is low. Owing to the shear stress acting, which is instigated by the bending fretting fatigue stress,
Chowdhury et al. 21

Figure 30. TEM images of microstructure of 316L steel in the crack initiation zone of bending fretting fatigue:
 a,max ¼ 509.3 MPa (Peng et al., 2014; Majzoobi and Minaii, 2013).

disengagement starts, and then single slip and cross-slip of separation occur. Furthermore, stacking
deficiency spreads. With the increase in neighborhood thickness of disengagement to a high level, the
separation slipping is limited, so the twinning framework is gradually developed. With the plastic
disfigurement further accumulating, the trifurcate twin structure and the triangle zones are empha-
sized, and disengagements are less likely to appear as coincidence, the thickness of separation
quickly expands, and the stress fixation in the triangle zone of the lattices produces; finally, as the
stress fixation surpasses the quality of the material, the smaller scale crack cores start with the
connecting of the adjoining crack cores. As shown in Figure 30, the miniaturized scale cracks
start and propagate in a zig-zag manner.
Separation cell nucleation model for the BCC structure metals and several FCC-structured metals
with high-level stacking fault energy (for instance, aluminum compound). In the microstructure of
the first 7075 aluminum compound and structure steel, the separation thickness is low, and a sep-
aration cross-slip occurs, thereby facilitating fine scattered stages. As a result of shear stress, which is
affected by the bending fretting fatigue load, the thickness of separation expands and the disengage-
ments tangle around the fine accelerated stages, resulting in the start of the disengagement cell
disfigurement. Moreover, with the accumulation of plastic twisting in local area, genuine heap up
disengagement occurs because of the sticking activity of slip band, and then the separation cells
distort and disengagement conflicts with the mass of separation cell. With further accumulation of
plastic distortion, the thickness of disengagement increases and the separation cells twist enormously
and dynamically decrease the size, resulting in higher stress focus to start the fretting fatigue crack
nuclei at the limit of the mass of disengagement cells. Compared with the 7075 aluminum composite,
the LZ50 steel had lesser separation cell structure. The commonplace development of separation
setup for this case is shown in Figure 31.

Case of Ti–6Al–4V alloy (shoe and bolt type rig) (Ebara and Fujimura, 2006)
Figure 32 demonstrates a plainly visible perspective of the samples after fretting fatigue tests.
Fatigue crack initiated at the notched area where stress concentrates with the exception of on
contact region in the middle of the sample and shoe as shown in Figure 32(a), while Figure 32(b)
demonstrates fretting fatigue failure at the contact territory. The cocoa powder with dark shading
was observed on the fretting zone approximately 1 mm wide. Figures 33 and 34 show SEM
22 International Journal of Damage Mechanics 0(0)

Figure 31. TEM images of microstructure of 7075 alloy in the crack initiation zone of bending fretting fatigue:
 a,max ¼ 306.5 MPa (Peng et al., 2014; Majzoobi and Minaii, 2013).

Figure 32. Macroscopic view of samples after fretting fatigue test. (a) Fatigue failure, contact stress: 1075.8 MPa,
plane-bending stress: 398.1 MPa, number of cycle for failure: 7.7  106 cycles. (b) Fretting fatigue failure contact stress.
1394.5 MPa, Plane-bending stress. 393.2 MPa, number of cycles for failure. 5.9  105 cycles (Ebara et al., 2006; Nowell
et al., 2006).

perception consequences of surface morphologies of the sample surface of cocoa powder with dark
shading observed on the fretted zone approximately 1 mm wide. Figures 33 and 34 show SEM
perception consequences of surface morphologies of the sample surface of the fretting fatigue-
failed samples. Striation (Figures 33(b) and 34(b)) and fretting debris (Figures 33(c) and 34(c))
were most regularly viewed at the contact zone between the sample and shoe. Pits were likewise
observed, as shown in Figures 33(d) and 34(d). The pit starts when fretting debris was expelled from
the surface striation framed because of the contact slip development. These morphologies were
compared with the surfaces of fretting fatigue-failed samples disregarding the fretting fatigue-failure
number of cycles. Fretting fatigue crack surfaces are shown in Figures 35 and 36. The fretting fatigue
crack started from the surface (Figures 35(a) and 36(a)). The fretting fatigue crack start mode was a
microstructure-subordinate transgranular crack, Figures 35(b) and 36(b), while striation was
Chowdhury et al. 23

Figure 33. Surface morphology of the fretting fatigue failure contact stress: 1394.5 MPa, plane-bending stress:
393.2 MPa, number of cycles for failure: 5.9  105 cycles (Ebara et al., 2006; Nowell et al., 2006).

overwhelmingly observed on fretting fatigue crack propagating zone (Figures 35(d) and 36(d)).
These surface morphologies were observed in all fretting crack surfaces regardless of the number
of cycles of fretting fatigue to failure. The stress ratio R (the minimum to maximum stress) value was
0.05 for the fretting fatigue test.

Case of SUS 304 in H2 environment (bar-spring-tightened test rig) (Kubota et al., 2011; Peng
et al., 2011)
Figure 37 presents the morphology of the fretting damage delivered in every environment. The
observed ranges were in the region of the contact edge. After the contact surface of the sample
tried in air was secured with oxidized fretting wear particles, the fretted surface was observed after
eliminating oxidized fretting wear particles by light buffing. Interestingly, small fretting wear par-
ticles in the hydrogen gas were observed. In hydrogen gas, a manufacturing plant roof-like sample or
chevron design whose edges are perpendicular to the slip path was observed. This observation was
typical for the fretting damage created in the hydrogen gas. To illustrate the arrangement system of
the fretting damage in hydrogen gas, coordinating of the contact surfaces between the sample and
contact pad was performed.
Figure 38 shows images of the contact surfaces of the sample and contact pad and surface profiles
measured utilizing a confocal magnifying instrument. The area of observations was near the contact
24 International Journal of Damage Mechanics 0(0)

Figure 34. Surface morphology of the fretting fatigue failure. Contact stress: 1394.5 MPa, plane-bending stress:
353.0 MPa, number of cycles for failure: 1.41  106 cycles (Ebara et al., 2006; Nowell et al., 2006).

edge. The surface profiles were measured on the softened lines, as shown in the above images.
The surface deformations of the sample and the contact pad matched well each other. Thus, the
deformation of the contact surfaces seen in the sample tried in the hydrogen gas would be created by
exchange of the material. In vacuum and nitrogen gas, the fretting wear component transformed
from an oxidation overwhelming procedure to a bond predominant procedure was considered.
Given that no oxidized fretting wear particles are delivered in hydrogen gas, bond is viewed as
predominant, similar to the fretting in a vacuum and nitrogen gas (Bethune and Waterhouse, 1968;
Chen and Wei, 2014). The range of the damaged parts appeared to be little contrary to the ostensible
contact region. On the off chance that the contact load and rubbing power are transmitted through
the followed spots, that the stress state in the region of the followed range is assumed to be excep-
tionally serious (Cortese et al., 2014; Xue and Yang, 2014).

Case of Cr–CrN coated AL7075-T6 (simple rotary bending test rig) (Zalnezhad et al., 2013)
Figure 39(a) and (b) shows typical illustrations from cross-sectional perspectives of fractured
uncoated and Cr–CrN-covered samples under fretting fatigue test. The cracked surfaces comprise
of two regions: the fretting areas caused by bending of fretting pads and a tensile-region-made stress.
A typical SEM micrograph of a CrN-covered sample with bond of 2200 mN after breaking under
fretting fatigue at a stress of 200 MPa is shown in Figure 40. This figure shows the CrN film covering
the AL7075-T6 compound which goes about as antifriction under the fretting pads.
Chowdhury et al. 25

Figure 35. Fretting fatigue surface. contact stress. 1394.5 MPa, plane-bending stress: 393.2 MPa, number of cycle for
failure: 5.9  105 cycle, (a) initiation, (b) 0.5 mm from initiation, (c) 1 mm from initiation, (d) 3 mm from initiation
(Ebara et al., 2006; Nowell et al., 2006).

Case of fatigue resistance of Al7075 (Moore test rig) (Majzoobi and Minaii, 2013; Luo, 2015)
Fracture surface and fretting zone of the samples were analyzed utilizing OM. A common place slip–
stick contact area where fretting occurs is shown in Figure 41. The stick area is caused by the high
contact pressure which keeps the pads from sliding in the contact area. The crack surface of the
sample D3 is outlined in Figure 42. The sample was subjected to fretting fatigue at the bending stress
of 220 MPa. The surface comprises of three distinct regions. (1) A fretting region where cracks start.
Two fretting regions found in the figure compare to the area of the contact ranges between the pads
and the sample. (2) A fatigue zone where crack propagates and has a shoreline mark appearance. (3)
An elastic range where crack occurs. As the crack propagates, the material loses its resistance to
crack, and when the fatigue range broadens sufficiently and the sample is weakened enough, the
sample cracks by strain. As shown in Figure 42, the tensile area is extensively bigger than the fretting
and fatigue regions. The augmentation of tensile area relies on the applied bending stress. The crack
surfaces of two received nonrolled samples under the load (a) 189.82 MPa and (b) 99.84 MPa are
shown in Figure 43. The ductile zone is considerably lesser for the bending load of 99.84 MPa than
that for the bending load of 189.82 MPa. The crack surfaces of two samples under bending loads of
207 and 369 MPa are delineated, as shown in Figure 44. These two loads relate to the high- and low-
cycle fatigue regimes. The crack surface that relates to the high-cycle fatigue regime is less rough
than that of the low-cycle fatigue regime.
26 International Journal of Damage Mechanics 0(0)

Figure 36. Fretting fatigue fracture surface. Contact stress: 1394.5 MPa, plane-bending stress: 353.0 MPa, number of
cycles for failure. 1.41  106 cycles, (a) initiation, (b) 0.5 mm from initiation, (c) 1 mm from initiation, and (d) 3 mm
from initiation (Ebara et al., 2006; Nowell et al., 2006).

Figure 37. Morphology of fretting damage ( a,max ¼ 200 MPa, N ¼ 105): (a) in air and (b) in H2 gas (Kubota et al.,
2011; Peng et al., 2011).
Chowdhury et al. 27

Figure 38. Matching of contact surfaces fretted in H2. The observed area was in the vicinity of the contact edge:
photographs of (a) sample, (b) contact pad, and (c) superposition of the surface profiles (Kubota et al., 2011; Peng
et al., 2011).

Figure 39. A typical sample of microscopic image from cross-sectional view of fractured samples under fretting
fatigue test (a) uncoated and (b) Cr–CrN-coated (Zalnezhad et al., 2013).
28 International Journal of Damage Mechanics 0(0)

Figure 40. Typically SEM micrograph of CrN-coated sample with adhesion of 2200 mN after fracture at stress of
200 MPa (Zalnezhad et al., 2013).

Figure 41. A slip–stick condition in fretting area (Majzoobi and Minaii, 2013; Zalnezhad et al., 2014).

Figure 42. Fracture surface of sample D3 (Majzoobi and Minaii, 2013; Zalnezhad et al., 2014).
Chowdhury et al. 29

Figure 43. A comparison between the fracture surfaces of two as received nonrolled samples under the load (a)
189.82 MPa and (b) 99.84 MPa (Majzoobi and Minaii, 2013; Zalnezhad et al., 2014).

Figure 44. Fracture surfaces of two samples tested under a low and b high-cycle fatigue conditions (Majzoobi and
Minaii, 2013; Luo, 2015).

Case of regular steel, Swedish standard 2140 (shrink-fitted joint) (Gutkin et al., 2008;
Zalnezhad et al., 2014)
In Figure 45(a) and (b), the advancement of the crack propagation is expected. At the highest point
of Figure 45(a) and in the growth in Figure 45(b) the initial crack can be observed. The semi-circular
30 International Journal of Damage Mechanics 0(0)

Figure 45. (a) Cross-section of a failed fretting sample after shrink-fit contact and rotating bending load. (b) Close-
up view on the initial fretting fatigue crack (Gutkin et al., 2008; Zalnezhad et al., 2014).

shape is noted (minor semi axis, into the pole, a ¼ 0.2 mm and real semi-pivot, along the surface,
b ¼ 0.3 mm). At approximately one-fourth from the highest point of the segment, a distinction in
surface difference indicates a temporary area of the crack front during the development, in which the
semicircular shape is conserved by all accounts rationed. Lastly, the special zone with a rougher
surface that covers the lower part of the cylinder corresponds to the last loaded region. In this
manner, the last crack size could be considered the furthest area of this region. Few helpful compo-
nents were identified based on the preceding discussion. A starting crack size was determined, which
was utilized for the numerical reproduction of the fatigue crack development life. The crack front
could be approximated by a semi-circular shape and the break developed in a plane opposite to the
length axle of the shaft at the length position of the contact edge or up to 1 mm inside the contact.

Chinese carbon railway axle RD2 (complex type test rig) (Song et al., 2014; Majzoobi and
Minaii, 2013)
The propagation of crack in Figure 46(a) for the real axle can be separated into two stages. At depth of
less than 150 mm, the crack propagated along a slope edge of approximately 27 to the radial direction
of the axle. The slope point of the crack slowly turned to a lower point of 21 with the increase in
depth. The propagation depth of the fretting fatigue crack is observed at approximately 320 mm.
Clearly, the first propagation stage was overwhelmed by the contact stresses of interference fit.
Nevertheless, the fatigue crack propagation procedure was controlled by the combination of contact
stresses and bulk stresses when the depth was 150 mm. However, three stages for crack propagation
process in 46(b) exist. Propagation proceeded along 25 from depth of less than 60 mm, after which the
crack turned to 21 . As depth reaches 150 mm, contact stress action disappeared.
Macroscopic views of fracture surface and contact edge are shown in Figure 47(a) and (b). The
surface can be divided into three regions according to the nature of crack growth. Figure 47(c) and
(d) shows high-magnification local morphologies of fracture surface. Crack initiation zone indicates
rotary bending fretting crack was in multisource feature, which shows good agreement with real axle
analysis. Multisource behavior is related to low cycle fatigue conditions that are controlled by plastic
deformation. Importantly, fretting wear is combined with local high contact stresses for rotary
bending fretting fatigue condition.
Chowdhury et al. 31

Figure 46. Scanning electron microscope image of the fretting fatigue crack observed in the inner fretting scars (a) the real
axle sample and (b) the small-scale axle sample under bending stress of 144 MPa, 4.5  106 cycles (Song et al., 2014; Majzoobi
and Minaii, 2013).

Figure 47. Fracture surface of small-scale axle sample under bending stress of 144 MPa (1800 r/min, 4.5  106
cycles) (a) Macro-morphology of cross section, (b) Macro-morphology of fracture surface, and (c) and (d) scanning
electron microscope image of the crack initiation area (Song et al., 2014; Majzoobi and Minaii, 2013).
32 International Journal of Damage Mechanics 0(0)

There are several studies are conducted in recent periods by different researchers (Das and
Pradhan, 2014; Genet et al., 2013; Vu et al., 2014; Wahl et al., 2013; Yao and Basaran, 2013)
which described the trends of the materials failure under different concerned mechanisms. These
sources of researches can be used as a basis for future studies.

Conclusions
Experimental details of different bending fretting fatigue phenomena were reviewed. Different rigs,
including reciprocating and rotary rigs, exhibited specific characteristics according to the pre-
assumed desired results for optimizing loading value, control, fretting pad type, and other operating
parameters. Findings of the above discussions can be summarized as follows:

(i) In the case of servo-hydraulic test rig on different samples, the S–N curve takes a shape of Greek
letter ‘‘"‘‘ with three regimes. For the 7075 Al alloy, the S–N curve displayed a C-like curve. The
fretting contact zones of SUS 316L were secured by some oxidative debris in the seriously
damaged zone. With the increase in the bending fatigue load, the microscopic cracks created
were in the shape of a zig-zag.
(ii) Shoe- and bolt-type rig using Ti–6Al–4V alloy shows fretting fatigue crack initiates over the
contact zone at higher stress. Fatigue crack initiation mode was structure-dependent transgra-
nular, and fretting fatigue crack propagation mode was striated.
(iii) The SUS 304 in H2 environment bar spring tightened test rig, adhesion prevails over fretting
damage in the H2 gas. The major crack leading to sample failure began from one of the little
cracks that exuded at the adhered spots.
(iv) In the case of Cr–CrN-coated AL7075-T6 sample in rotary bending test rig, coating increases
fretting fatigue life compared with that of an uncoated sample.
(v) In fatigue resistance of Al7075 Moore test rig, the fatigue life of Al7075 was reduced by 95%
because of fretting. Deep rolling is significantly more effective in high-cycle fatigue regime,
whereas its effect in low-cycle fatigue regime is not as significant.
(vi) For complex type test rig using Chinese carbon railway axle RD2, fretting fatigue damage scars
can be divided into three regimes where cracks were found in inner scars for both the real and
small-scale axles. Cracks initiated at the subsurface and propagated along an inclined angle.

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or pub-
lication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication of this
article.

References
Bethune B and Waterhouse R (1968) Adhesion of metal surfaces under fretting conditions I. Like metals in
contact. Wear 12(4): 289–296.
Chen L and Wei Y (2014) Overall mechanical behavior of nanocrystalline materials accompanied by damage
evolution on grain boundaries. International Journal of Damage Mechanics 23(1): 25–42.
Chowdhury et al. 33
Chuanfu YWZWD and Zenye Z (1992) Study on spectrum load fatigue behaviors, fracture mechanism and
fractography of two ultra-high strengh steels. Journal of Materials Engineering 2: 004.
Cortese L, et al. (2014) Prediction of ductile failure in materials for onshore and offshore pipeline applications.
International Journal of Damage Mechanics 23(1): 104–123.
Dao NH and Vu MN (2016) Load sequence effects on the fatigue crack growth in a cylinder subjected to
combined rotary bending moment and axial force loads. Theoretical and Applied Fracture Mechanics 82:
117–124.
Das RR and Pradhan B (2014) Delamination damage analysis of laminated bonded tubular single lap joint
made of fiber-reinforced polymer composite. International Journal of Damage Mechanics 23(6): 772–790.
Ding J, et al. (2014) Finite element analysis on bending fretting fatigue of 316L stainless steel considering
ratchetting and cyclic hardening. International Journal of Mechanical Sciences 86: 26–33.
Ebara R and Fujimura M (2006) Fretting fatigue behaviour of Ti–6Al–4V alloy under plane bending stress and
contact stress. Tribology International 39(10): 1181–1186.
Foletti S, Beretta S and Gurer G (2016) Defect acceptability under full-scale fretting fatigue tests for railway
axles. International Journal of Fatigue 86: 34–43.
Gaudin C and Feaugas X (2004) Cyclic creep process in AISI 316L stainless steel in terms of dislocation
patterns and internal stresses. Acta Materialia 52(10): 3097–3110.
Genet M, Marcin L and Ladeveze P (2014) On structural computations until fracture based on an anisotropic
and unilateral damage theory. International Journal of Damage Mechanics 23(4): 483–506, DOI: 10.1177/
1056789513500295.
Goidescu C, et al. (2015) Anisotropic unilateral damage with initial orthotropy: A micromechanics-based
approach. International Journal of Damage Mechanics 24(3): 313–337.
Gosar A and Nagode M (2015) Dissipated energy-based fatigue lifetime calculation under multiaxial plastic
thermo-mechanical loading. International Journal of Damage Mechanics 24(1): 41–58.
Gutkin R and Alfredsson B (2008) Growth of fretting fatigue cracks in a shrink-fitted joint subjected to rotating
bending. Engineering Failure Analysis 15(5): 582–596.
Haidyrah AS, Newkirk JW and Castaño CH (2016) Weibull statistical analysis of Krouse type bending fatigue
of nuclear materials. Journal of Nuclear Materials 470: 244–250.
Hattori T, et al. (1988) Fretting fatigue analysis using fracture mechanics. JSME International Journal. Ser. 1,
Solid Mechanics, Strength of Materials 31: 100–107.
Hills D and Nowell D (1994) Mechanics of Fretting Fatigue. Solid Mechanics and its Applications. Vol. 30.
Dordrecht: Kluwer Academic Publishers.
Huang L, et al. (2016a) Fatigue and fretting of mixed metal self-piercing riveted joint. International Journal of
Fatigue 83: 230–239.
Huang Y, Wang Z and Zhou Q (2016b) Numerical studies on the surface effects caused by inhomogeneities on
torsional fretting. Tribology International 96: 202–216.
Hwang B, Kim T and Han SM (2016) Compression and tension bending fatigue behavior of Ag nanowire
network. Extreme Mechanics Letters. http://dx.doi.org/10.1016/j.eml.2016.02.011.
Iwabuchi A, Kayaba T and Kato K (1983) Effect of atmospheric pressure on friction and wear of 0.45% C steel
in fretting. Wear 91(3): 289–305.
Johnson K (1995) Contact mechanics and the wear of metals. Wear 190(2): 162–170.
Johnson K and Greenwood J (1997) An adhesion map for the contact of elastic spheres. Journal of Colloid and
Interface Science 192(2): 326–333.
Kamaya M (2016) Development of disc bending fatigue test technique for equi-biaxial loading. International
Journal of Fatigue 82: 561–571.
King RN and Lindley TC (1981) Fretting fatigue in a 3/2 Ni–Cr–Mo–V rotor steel. In: Francosis D (ed.) In:
Proceedings of ICF5. Oxford: Pergamon Press, p. 631.
Kubota M, et al. (2011) Mechanism of reduction of fretting fatigue limit caused by hydrogen gas in SUS304
austenitic stainless steel. Tribology International 44(11): 1495–1502.
Lee H and Mall S (2006) Investigation into effects and interaction of various fretting fatigue variables under
slip-controlled mode. Tribology International 39(10): 1213–1219.
34 International Journal of Damage Mechanics 0(0)

Li K, et al. (2016) Fretting fatigue characteristic of Ti–6Al–4V strengthened by wet peening. International
Journal of Fatigue 85: 65–69.
Liu W, et al. (2013) The performances of a thermally sprayed Fe/Ni composite coating to resist fretting fatigue
under rotational bending loads. Surface and Coatings Technology 217: 58–63.
Luo RK (2015) Mullins damage effect on rubber products with residual strain. International Journal of Damage
Mechanics 24(2): 153–167.
Ma Y, et al. (2015) The bending fatigue performance of cement-stabilized aggregate reinforced with polypro-
pylene filament fiber. Construction and Building Materials 83: 230–236.
Majzoobi G and Minaii K (2013) An investigation into the effect of contact geometry on the rotary bending
fretting fatigue life of Al 7075-T6. Proceedings of the Institution of Mechanical Engineers, Part J: Journal of
Engineering Tribology 227(11): 1285–1296.
Mayama T, Sasaki K and Kuroda M (2008) Quantitative evaluations for strain amplitude dependent organ-
ization of dislocation structures due to cyclic plasticity in austenitic stainless steel 316L. Acta Materialia
56(12): 2735–2743.
Mdzimba P (2006) Passningsrost vid pressförband. Master Thesis. Stockholm: KTH Solid Mechanics, Sweden
[in Swedish].
Neu RW, Pape JA and Swalla DR (2000) In: Hoeppner DW, Chandrasekaran V and Elliott CB (eds) Fretting
Fatigue: Current Technology and Practices, ASTM STP 1367. West Conshohocken, PA: ASTM, 2000, pp.
369–388.
Nowell D, Dini D and Hills D (2006) Recent developments in the understanding of fretting fatigue. Engineering
Fracture Mechanics 73(2): 207–222.
Peng J, et al. (2011) An experimental study on bending fretting fatigue characteristics of 316L austenitic
stainless steel. Tribology International 44(11): 1417–1426.
Peng J, et al. (2014) On the damage mechanisms of bending fretting fatigue. Tribology International 76:
133–141.
Peng J-F, et al. (2013) Study on bending fretting fatigue damages of 7075 aluminum alloy. Tribology
International 59: 38–46.
Savaria V, Bridier F and Bocher P (2016) Predicting the effects of material properties gradient and residual
stresses on the bending fatigue strength of induction hardened aeronautical gears. International Journal of
Fatigue 85: 70–84.
Shi L, et al. (2016) An investigation of fretting fatigue in a circular arc dovetail assembly. International Journal
of Fatigue 82: 226–237.
Song C, et al. (2014) An investigation on rotatory bending fretting fatigue damage of railway axles. Fatigue &
Fracture of Engineering Materials & Structures 37(1): 72–84.
Vu QH, Halm D and Nadot Y (2014) High cycle fatigue of 1045 steel under complex loading: Mechanisms map
and damage modelling. International Journal of Damage Mechanics 23(3): 377–410.
Wahl L, et al. (2013) Fatigue in the core of aluminum honeycomb panels: Lifetime prediction compared with
fatigue tests. International Journal of Damage Mechanics 23(5): 661–683.
Winkler J, Georgakis CT and Fischer G (2015) Fretting fatigue behavior of high-strength steel monostrands
under bending load. International Journal of Fatigue 70: 13–23.
Xue X and Yang X (2014) A damage model for concrete under cyclic actions. International Journal of Damage
Mechanics 23(2): 155–177.
Yaguchi H, et al. (2001) Fatigue-damage evaluation in aluminum heat-transfer tubes by measuring dislocation
cell-wall thickness. Materials Science and Engineering: A 315(1): 189–194.
Yao W and Basaran C (2013) Damage mechanics of electromigration and thermomigration in lead-free solder
alloys under alternating current: An experimental study. International Journal of Damage Mechanics 23(2).
Yang Y, Lam C and Kou K (2014) Fatigue life analysis of internal circumferential crack of tubular structures.
International Journal of Damage Mechanics 24(5).
Yuan R, et al. (2014) A nonlinear fatigue damage accumulation model considering strength degradation and its
applications to fatigue reliability analysis. International Journal of Damage Mechanics 24(5).
Chowdhury et al. 35
Zalnezhad E, Sarhan AA and Hamdi M (2013) Fretting fatigue life evaluation of multilayer Cr–CrN-coated
Al7075-T6 with higher adhesion strength—fuzzy logic approach. The International Journal of Advanced
Manufacturing Technology 69(5–8): 1153–1164.
Zalnezhad E, Sarhan AA and Jahanshahi P (2014) A new fretting fatigue testing machine design, utilizing
rotating–bending principle approach. The International Journal of Advanced Manufacturing Technology
70(9–12): 2211–2219.
Zhou Z and Zhu M (2004) Composite Fretting Wear. Shanghai: Shanghai Jiao Tong University.
Zhu H, et al. (2016) A two-dimensional micromechanical damage–healing model on microcrack-induced
damage for microcapsule-enabled self-healing cementitious composites under compressive loading.
International Journal of Damage Mechanics 25(5).

View publication stats

You might also like