Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Wear 444-445 (2020) 203120

Contents lists available at ScienceDirect

Wear
journal homepage: http://www.elsevier.com/locate/wear

Effects of carbon content and hardness on rolling contact fatigue resistance


in heavily loaded pearlitic rail steels
Masaharu Ueda a, *, Kenji Matsuda b
a
Nippon Steel Corporation, 1-1 Tobihata-cho, Tobata-ku, Kitakyushu-shi, Fukuoka, 804-8501, Japan
b
Kyushu Institute of Technology, 1-1 Sensui-cho, Tobata-ku, Kitakyushu-shi, Fukuoka, 804-8550, Japan

A R T I C L E I N F O A B S T R A C T

Keywords: To reproduce the rolling contact fatigue (RCF) damage of rail steel used in heavy haul railways and to clarify its
Rolling contact fatigue possible mechanism, the RCF characteristics of pearlitic steel with different carbon content and initial hardness
Rail values are evaluated using a two-disk-type machine. The results indicate that the number of spalls and the crack
Pearlitic steel
depth decrease with increasing carbon content of pearlitic steel even when the initial hardness is almost the same
Carbon content
Wear
level. Furthermore, an increase in the carbon content increases the hardness of the rolling contact surface,
Plastic deformation resulting in the suppression of the plastic flow development. Results also indicate that the crack inclination angle
Crack varies depending on the carbon content. According to these results, a possible mechanism for the suppression of
fatigue crack propagation in higher carbon pearlitic steel is proposed on the basis of the fracture mechanics
approach.

1. Introduction without increasing the cost, inhibition of RCF damage of rails is required
[7,8].
In heavy haul railways transporting goods, such as natural resources, Numerous studies regarding the RCF damage of rails induced by
including ores and cereals, more heavily loaded freight is actively being cumulative plastic deformation have been conducted by various re­
promoted for more efficient transportation. As a result, the wheel load of searchers from different viewpoints. From the viewpoint of RCF damage
freight cars is significantly increasing, making the operating environ­ prediction, multi-axial fatigue evaluation models have been applied to
ment of rails more severe. Thus, the rolling contact fatigue (RCF) evaluate the crack initiation characteristics under the complex contact
damage is increasing [1]. conditions between rails and wheels [9]. On the basis of this analysis,
An example of the RCF damage generated in heavy haul railways is the service life of rails has been predicted [10,11]. Moreover, from the
shown in Fig. 1. Contact pressure and tangential traction due to the viewpoint of suppressing RCF damage, optimization of lubrication and
rolling of the wheel are exerted on the contact surface between the rail improvement of rail steels have been studied. In terms of lubrication, to
and wheel. In particular, at the gauge corner (G.C.) portion of a rail, reduce the tangential force generated at the contact portion between the
slippage occurs due to its contact with the throat part of the wheel, rail and wheel, which affects the occurrence of RCF damage, lubricants
resulting in plastic flow development due to tangential force beneath the and their control methods have been developed, thus improving the
rolling contact surface. In addition, cracks initiate at plastic flow por­ service life of rails [12–14]. Furthermore, from the viewpoint of
tions due to repeated contact, leading to RCF damage, also referred to as improving the characteristics of rail steel, their correlation with the
spalling, which is caused by crack propagation [2–6]. structure [15,16] and hardness [17–19] of the steel has been discussed;
As the typical RCF damage of this type involves crack propagation, it has been shown that increasing the hardness of pearlitic steel without
they could induce rail breakage in addition to spalling. Therefore, pro-eutectoid ferrite structure effectively inhibits RCF damage. As a
maintenance control of tracks via regular preventive grinding is result, various high-strength (hardness) rails with the pearlitic structure
required to ensure operating safety of the rail. When the RCF damage have been developed [20–22] and have contributed to improved service
increases, the frequency of preventive grinding increases, which in turn life of rails [23].
increases the maintenance control cost. To improve the safety of tracks The pearlitic structure is a lamellar structure of the ferrite phase and

* Corresponding author.
E-mail address: ueda.ph4.masaharu@jp.nipponsteel.com (M. Ueda).

https://doi.org/10.1016/j.wear.2019.203120
Received 31 May 2019; Received in revised form 28 October 2019; Accepted 8 November 2019
Available online 15 November 2019
0043-1648/© 2019 Elsevier B.V. All rights reserved.
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

cementite phase. It is well known that the space between the ferrite
phase and cementite phase is referred to as the lamellar spacing, and the
hardness of pearlitic structure increases when the lamellar spacing is
refined [24]. RCF damage caused by plastic deformation is inhibited by
refining the lamellar spacing, thereby improving the service life of the
rail [17–19]. In addition, an increase in the carbon content of pearlitic
steel allows a volume fraction of the cementite phase in pearlitic la­
mellas to increase [25]. It has been reported that an increase of the
volume fraction of the cementite phase in pearlitic lamellas could in­
crease hardness of the rolling contact surface, which would improve
wear resistance [26]. Accordingly, an increase in the carbon content in
pearlitic steel could inhibit the RCF damage caused by plastic defor­
mation formed in heavy haul railways and thus improve the RCF
resistance.
The purpose of this study is to clarify the effects of carbon content
and the initial hardness of pearlitic steel on its RCF resistance under
heavy rolling/sliding contact condition. RCF damage of pearlitic steels
having different carbon content and initial hardness values is compared
using a two-disk-type machine model of the rail/wheel contact in actual
heavy haul railways. In addition, a possible mechanism for the occur­
rence of differences in the degree of spalling depending on the differ­
ences in carbon content of pearlitic steel has been studied on the basis of
the fracture mechanics approach.

2. Experimental procedure
Fig. 1. Example of a damaged rail in a heavy haul railway: (a) View of a gauge
corner with developed rolling contact fatigue defects, (b) Cross-sectional
microstructure of rolling contact fatigue cracking. 2.1. Testing machine

The image and the schematic outline of a two-disk-type testing ma­


chine are shown in Fig. 2. The wheel disk and rail disk can be separately

Fig. 2. Image and schematic illustration of the testing machine for evaluating rolling contact fatigue resistance.

2
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

driven by motor via a belt. In addition, the load in the radial direction
and the load in the thrust direction can be separately provided by a
hydraulic system. The angle of the wheel axis can be changed to provide
a certain inclination to the rail axis, i.e., the angle of attack.

2.2. Specimens

The carbon content, initial hardness and microstructure of the


specimens are listed in Table 1. Here the hardness was measured by the
Vickers hardness test with a load of 10 kgf. The rail specimens were
steels containing carbon levels of 0.8 and 1.0 mass% to which Si, Mn,
and Cr alloys are added. Hereafter, the rail specimens with a carbon
content of 0.8 mass% are referred to as 0.8mass%C steel, and those with
a carbon content of 1.0 mass% are referred to as 1.0mass%C steel. In
addition, the wheel specimen was steel containing a carbon content of
0.8 mass% to which Si, Mn, and Cr alloys are added.
The specimen materials for rail and wheel were obtained via vacuum
melting with electrolytic iron and alloy iron. Steel ingots were produced
after the chemical composition control in vacuum melting and forged to
diameters 220–240 mm by reheating at 1250 � C. To prepare the three
types of rail specimens with different initial hardness values of the
pearlitic structure and to prevent the formation of a pro-eutectoid
cementite structure, the forged materials were roughly machined into
the required size of the specimens, reheated via induction at 950

C 1000 � C, and then subjected to accelerated cooling. The initial
hardness was controlled to be at three levels ranging from 350 to 415 HV
for 0.8mass%C steel and 1.0mass%C steel respectively.
For the wheel specimen, the forged materials were roughly machined
into the required size of the specimens, reheated via induction at 950 � C,
and then, subjected to accelerated cooling; the initial hardness was
maintained at 350 HV.
The sectional profiles of the rail and the wheel specimens are shown
in Fig. 3. The surface profile of each rail and wheel specimen was scaled
down to one quarter of a full-scale rail and wheel used in heavy haul
railways. The diameter at the central portion of the width of the rail
specimen was the same as that of the wheel specimen and was 200 mm.
The rail specimen is brought into contact with the wheel specimen at
two main portions; the rail head and the G.C. portion. Slippage will
occur at the G.C. portion where the throat portion of the wheel specimen Fig. 3. Cross-sectional dimensions of rolling contact fatigue test specimens: (a)
contacts with the rail, because of the differences in the rotation radius Rail specimen, (b) Wheel specimen.
depending on the contact portion [27]. The rotation radius of the throat
portion is larger than that of the G.C. portion even though that of the
tread portion of the wheel specimen is similar to that of the rail head Table 2
portion; therefore, during the test, surface traction was provided in the Testing conditions.
same direction as the rolling direction (opposite to the direction of load Loading (kN) Radial: 17.7 Thrust: 11.8
movement). Angle of attack (� ) 0.5
Initial contact pressure (MPa) 2000 (G.C. portion)
The total number of rolling contact 550 � 103
2.3. Testing condition cycles
Lubrication and its duration of cycles Dry (Non-lubrication)/0 � 103 to 20 � 103
The testing conditions are shown in Table 2. The wheel disk was Water (5 cc/min)/20 � 103 to 320 � 103.
Repeated dry (1.0 min) and water (0.5 min)
driven by the motor, whereas the rail disk was set to rotate freely and /320 � 103 to 550 � 103
then driven by the wheel disk. The loading condition to reproduce the Rotating speed (rpm)
contact condition along a curved track (with a radius of curvature of Wheel 100 (Dry), 200 (Water, Repeated dry and
200–300 m) in the heavy haul railways having a wheel load of water)
Rail Driven by wheel
approximately 200 kN was assumed to be a radial load of 17.7 kN and
thrust load of 11.8 kN. In addition, the wheel disk was provided with an
angle of attack of 0.5� . The initial contact pressure of the G.C. portion 2000 MPa (please see Section 4.1 for details).
was calculated using finite element method (FEM) and was estimated as The total number of rolling contact cycles was set to 550 � 103. The
lubrication condition was as follows. The initial period up to 20 � 103
Table 1 cycles was under dry condition. The second period from 20 � 103 to 320
Chemical composition, initial hardness, and microstructure of test specimens. � 103 cycles was under water lubrication condition. In the last period
Specimen Carbon content (mass%) Initial hardness (HV10) Microstructure from 320 � 103 to 550 � 103 cycles, dry condition period for 1.0 min and
Rail 0.81 350, 375, 400 Pearlite
water lubrication condition period for 0.5 min were alternately
1.00 350, 385, 415 Pearlite repeated. Under lubrication condition, water was dropped between two
Wheel 0.80 350 Pearlite disks at 5 cc/min.

3
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

The aim of these lubrication conditions was as follows: The initial dry direction and etched by nital before the microstructures and cracks were
running was to perform running in. However, after the initial dry observed using an optical microscope. Then, the length of plastic flow
running, the surface roughness increased. The subsequent wet running and crack inclination angles were measured.
provided a relatively smooth contact surface similar to that of the actual The hardness distribution in the depth direction of the cross-section
rail surface and simultaneously generated micro-cracks beneath the beneath the rolling contact surface was also measured by the micro-
rolling contact surface. In addition, to reproduce the RCF damage caused Vickers hardness test with a load of 0.025 kgf. In addition, to evaluate
by plastic deformation, the repeated dry and water lubrication condi­ the wear characteristics, the cross-sectional head shape of the rail
tions provided the rolling contact surface with strong tangential trac­ specimens after the test was measured by using a 3D coordinate
tion, which facilitated plastic flow during dry running and allowed measuring machine (Mitutoyo, M544) with contact probe. Furthermore,
water to penetrate into the inside of the cracks during wet running, the shape after the test and the initial shape were superimposed with
promoting crack propagation [28,29]. reference to the side opposite to the G.C. part where wear was not
Moreover, to investigate the variation of the microstructure and the generated, and the difference of the cross-sectional area between before
crack propagation beneath the rolling contact surface under the testing, and after the test was taken as the amount of wear.
two sets of additional tests were conducted on the specimens of 0.8mass Note that two sets of the specimens were used for each test at each
%C steel with the initial hardness of 400 HV and 1.0mass% C steel with value of hardness and each number of cycles. The degree of spalling on
the initial hardness of 415 HV. One test was terminated after 320 � 103 the rolling contact surfaces was visually inspected on the quarter
cycles, corresponding to the end of the water lubrication condition and circumference of each specimen. By contrast, microstructure inspection
another after 420 � 103 cycles, corresponding to the middle of the and hardness measurement of the cross-section were conducted at two
repeated dry and water lubrication conditions. observation sites of an arc length of 8 mm on the rolling contact surface
along the quarter circumference of each specimen. Wear amount mea­
surement was performed on one cross section in the quarter circumfer­
2.4. Evaluation
ence of each specimen.

Rolling contact surface of the rail specimens after the tests was
inspected visually. Furthermore, to evaluate the crack generation con­
dition and structural change beneath the rolling contact surface, G.C.
portions where cracks were dominantly generated were cut in the rolling

Fig. 4. Typical rolling contact surfaces of the rail specimens after the test:
(a) 0.8mass%C 350 HV, (b) 0.8mass%C 375 HV, (c) 0.8mass%C 400 HV, (d) 1.0mass%C 350 HV, (e) 1.0mass%C 380 HV, (f) 1.0mass%C 415 HV.

4
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

3. Experimental results carbon content was smaller than that of 0.8mass%C steel at the same
hardness level.
3.1. Effects of carbon content and initial hardness on the RCF resistance
3.1.3. Comparison of wear amount
3.1.1. Comparison of the rolling contact surface The relationship between the initial hardness and wear amount of
The rolling contact surfaces for each rail specimen after 550 � 103 the rail specimens is shown in Fig. 8. Here, the wear amount was eval­
cycles are shown in Fig. 4. Various sizes of pits or spalls were observed in uated using one cross-section for each rail specimen. Fig. 8 presents the
a scattered state in the rolling direction of the G.C., and the length of the results of every two specimens and their mean value for each kind of rail
maximum spall reached almost 10 mm in the rolling direction. The steel. Irrespective of the carbon content, the wear amount was corre­
number and the size of spalls were found to vary depending on the kind lated with the initial hardness of the specimens, that is, the wear amount
of rail steel. decreased as the initial hardness of the specimens increased. In addition,
To quantify the spalling occurrence status, the number of spalls of the wear amount of 1.0mass%C steel with higher initial hardness
length 2 mm or more in the rolling direction was counted at the G.C. decreased compared with that of 0.8mass%C steel at the same hardness
portion. The relationship between the initial hardness of the specimens level.
and number of spalls is shown in Fig. 5. In this figure, the results of two
specimens and their mean values are plotted for each kind of rail steel. 3.2. Effect of the rolling contact cycles on the growth of RCF damage
Irrespective of the carbon content of the specimens, the number of spalls
decreased as the initial hardness of the specimens increased. In addition, In the previous section, the RCF properties of pearlite steel were
with respect to the correlation with the carbon content of the specimens, evaluated, and it became clear that the RCF resistance of pearlite steel is
the number of spalls in 1.0mass%C steel with higher carbon content improved by the increase of carbon content. In this chapter, in order to
significantly decreased compared with that of 0.8mass%C steel at the clarify the mechanism of improvement in the RCF resistance with the
same hardness level. It was also apparent from Fig. 5 that increasing the increase in carbon content, two sets of additional tests were conducted
hardness of the 1.0mass%C steel resulted in a greater rate of decrease in on the specimens of 0.8mass%C steel with the initial hardness of 400 HV
the spalling damage than for the 0.8mass%C steel. and 1.0mass%C steel with the initial hardness of 415 HV. One test was
terminated after 320 � 103 cycles, corresponding to the end of the water
3.1.2. Comparison of cross-sectional aspect beneath the G.C. portion lubrication condition, and another after 420 � 103 cycles, corresponding
The typical cross-sectional aspects beneath the G.C. portion for each to the middle of the repeated dry and water lubrication conditions.
rail specimen are shown in Fig. 6. In all the specimens, the surface layer Here, the reason for comparing the results of specimen of 0.8mass%C
was displaced in the rolling direction, that is, plastic flow was identified steel with 400 HV and that of 1.0mass%C steel with 415 HV is as follows:
near the surface. Furthermore, many cracks initiated and propagated As shown in Fig. 5, the differences in the number of spalls between
along the plastic flow. However, some cracks branched and propagated 0.8mass%C steel and 1.0mass%C steel increase with increasing the
toward the upper surface. hardness of specimen. Therefore, the effect of the carbon content can be
Fig. 7 shows the relationship between the initial hardness and seen more clearly when the specimens with higher hardness are
maximum crack depth in the rail specimens. Note that the deepest crack compared. Furthermore, when comparing the number of spalls of
in one observation field with an arc length of 8.0 mm was measured as 1.0mass%C steel with 385 HV and that of 0.8mass%C steel with 400 HV,
the maximum crack depth, as shown in Fig. 6. Here two observation the same large/small relation is kept as that for 1.0mass%C steel with
fields were examined in every two specimens, that is, a total of four 415 HV and 0.8mass%C steel with 400 HV. Consequently, it is consid­
observation fields were examined for each kind of steel. In this figure, ered appropriate to compare the results of the specimens of 0.8mass%C
the mean value of the maximum crack depth and the dispersion range steel with 400 HV and 1.0mass%C steel with 415 HV in order to clarify
are marked. Irrespective of the carbon content, the maximum crack the effect of carbon content.
depth decreased with increasing initial hardness of the specimens. In
addition, the maximum crack depth of 1.0mass%C steel with higher 3.2.1. Variation of cross-sectional aspect
Comparison of the typical cross-sectional aspects beneath the G.C.
portion of the rail specimens at different rolling contact cycles is shown
in Fig. 9. In all specimens, plastic flow formation and crack generation
were identified beneath the rolling contact surface. At cycle number 320
� 103, corresponding to the end of water lubrication, irrespective of the
carbon content of the specimens, small plastic flow formation and small
crack generation were identified. At cycle number 420 � 103, which is
the middle of the repeated dry and water lubrication conditions, irre­
spective of the carbon content of the specimens, plastic flow developed,
and the crack depth increased compared with after the end of water
lubrication.
At final cycle number 550 � 103, in 0.8mass%C steel, plastic flow
developed further and the crack depth increased. In contrast, changes in
plastic flow and crack depth for 1.0mass%C steel were small compared
with those for 0.8mass%C steel.

3.2.2. Extent of plastic flow beneath the rolling contact surface


The relationship between the number of rolling contact cycles and
length of plastic flow is shown in Fig. 10. Here, the length of plastic flow
was determined by a method proposed by Eadie et al. [30], which was
based on the distance along the rolling direction from the start point of
Fig. 5. Relationship between the initial hardness and number of spalls with the plastic flow defined as the point at which the gradient of the plastic
lengths of 2 mm or more in the rolling direction formed on the rolling contact flow reaches 45� to the end point at which the plastic flow reaches the
surface of the rail specimens after the test. rolling contact surface as shown in Fig. 10. The maximum, minimum,

5
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 6. Typical cross-sectional aspects beneath the G.C. portion of the rail specimens after the test:
(a) 0.8mass%C 350 HV, (b) 0.8mass%C 375 HV, (c) 0.8mass%C 400 HV, (d) 1.0mass%C 350 HV, (e) 1.0mass%C 380 HV, (f) 1.0mass%C 415 HV.

Fig. 7. Relationship between the initial hardness and the maximum crack Fig. 8. Relationship between the initial hardness and wear amount of the rail
depth beneath the G.C. portion of the rail specimens after the test. specimens after the test.

6
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 9. Typical cross-sectional aspects beneath the G.C. portion of rail specimens after the test:
(a) 0.8mass%C 400 HV after 320 � 103 cycles, (b) 0.8mass%C 400 HV after 420 � 103 cycles, (c) 0.8mass%C 400 HV after 550 � 103 cycles, (d) 1.0mass%C 415 HV
after 320 � 103 cycles, (e) 1.0mass%C 415 HV after 420 � 103 cycles, (f) 1.0mass%C 415 HV after 550 � 103 cycles.

Fig. 10. Relationship between the number of rolling contact cycles and length of plastic flow formed beneath the G.C. portion of the rail specimen after the test.

and average values of typical 20 plastic flow lines observed for each between 0.8mass%C steel and 1.0mass%C steel. At cycle number 420 �
cycle and for each kind of steel are plotted in this figure. 103, which is in the middle of the repeated dry and water lubrication
At cycle number 320 � 103, corresponding to the end of water conditions, the plastic flow of 0.8mass%C steel significantly developed
lubrication, there was no clear difference in the length of plastic flow compared with that of 1.0mass%C steel. At the final cycle number 550 �

7
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

103, the length of plastic flow of 0.8mass%C steel increased furthermore, repeated dry and water lubrication conditions, the mean crack inclina­
whereas that of 1.0mass%C steel hardly changed. tion angle for 0.8mass%C steel and 1.0mass%C steel reduced to 21� and
30� , respectively. The mean crack inclination angle for 0.8mass%C steel
3.2.3. Variation of crack depth reduced further to 17� at the final cycle number of 550 � 103. Mean­
The relationship between the number of rolling contact cycles and while, the mean crack inclination angle for 1.0mass%C steel was 31� .
maximum crack depth is shown in Fig. 11. Note that the deepest crack in That is, during the repeated dry and water lubrication conditions from
one observation field with an arc length of 8.0 mm was measured as the cycle numbers 420 � 103 to 550 � 103, the plastic flow further devel­
maximum crack depth, as shown in Fig. 6. Here, the number of obser­ oped resulting in the reduction of the crack inclination angle for 0.8mass
vation fields under each cycle and each kind of steel is the same as in %C steel, whereas for 1.0mass%C steel, the crack inclination angle and
Section 3.1.2. In Fig. 11, the mean value of the maximum crack depth the length of plastic flow barely changed.
and the dispersion range are marked.
At cycle number 320 � 103, there was no clear difference in the 3.2.5. Variation of cross-sectional hardness
maximum crack depth between 0.8mass%C steel and 1.0mass%C steel. The micro-Vickers hardness distributions of the cross section cut
At cycle number 420 � 103, which is the middle of the repeated dry and along the G.C. of the rail specimen are shown in Fig. 13. Here, only the
water lubrication conditions, the maximum crack depth of 0.8mass%C average values of 20 measurements for each depth are plotted. Irre­
steel increased compared with that of 1.0mass%C steel. At final cycle spective of the carbon content of the specimens, the hardness beneath
number 550 � 103, the maximum crack depth of 0.8mass%C steel the rolling contact surface as well as the extent of hardness increasing
increased significantly, whereas that of 1.0mass%C steel showed a very area increased as the cycle number increased. At both cycle numbers
little increase from 420 � 103 cycle. That is, from 420 � 103 to 550 � 420 � 103 and 550 � 103 in the repeated dry and water lubrication
103 cycles, during the repeated dry and water lubrication conditions, conditions, the hardness beneath the rolling contact surface significantly
plastic flow further developed, resulting in the increase in the crack increased as the cycle number increased. In particular, the increase in
depth for 0.8mass%C steel; in contrast, in the case of 1.0mass%C steel, hardness was higher toward the surface.
the crack depth and the length of the plastic flow barely changed. The variations of the hardness and the hardness increase rate of the
surface layer as a function of the rolling contact cycles are shown in
3.2.4. Variation of crack inclination angle Fig. 14. Here, the hardness at the surface layer was the average value
The inclination angle of the cracks formed in the cross-section from the surface to a depth of 50 μm in the hardness distribution shown
beneath the G.C. portions was evaluated to quantify the crack genera­ in Fig. 13, which was considered to be the most susceptible to contact
tion characteristics. Here, the crack inclination angle for a curved crack, with the wheel. The hardness increase rate (HIR) is defined by:
such as that shown in Fig. 12, varies depending on its definition, which
will be discussed in detail below in Section 4.2. In this section, the crack HIR(%)¼(PH-IH)/IHx100 (1)
inclination angle was defined as an angle (β 200) formed between the where PH and IH are the post-test hardness and initial hardness,
rolling contact surface and line connecting the crack mouth and the respectively.
point on the crack at a depth of 200 μm as shown in Fig. 12. The reason The hardness of the surface layer increased almost linearly with the
for setting the crack depth to 200 μm was that the maximum depth of the rolling contact cycles for both 0.8mass%C steel and 1.0mass%C steel.
crack after the end of water lubrication was about 300 μm. The rela­ However, the 1.0mass%C steel with higher carbon content showed a
tionship between the number of rolling contact cycles and crack incli­ higher HIR than 0.8mass%C steel with lower carbon content. Conse­
nation angle is shown in Fig. 12. In this figure, 20 typical cracks having a quently, the hardness difference between these two steels increased
depth of 200 μm or more were selected and measured for each number of from the initial value of approximately 15 HV to approximately 50 HV at
cycles and for each kind of steel. the end of the test.
Irrespective of the carbon content, the mean crack inclination angle
was approximately 37� at cycle number 320 � 103, after the end of water 4. Discussion
lubrication. At cycle number 420 � 103, which is in the middle of the
As shown in Fig. 9, after the end of water lubrication at 320 � 103
cycles, micro-plastic flow and cracks were generated beneath the rolling
contact surface. However, no significant difference in the extent of
plastic flow beneath the rolling contact surface was identified between
rail specimens of 0.8mass%C steel and 1.0mass%C steel.
After the repeated dry and water lubrication conditions, the devel­
opment of plastic flow and cracks were identified for both 0.8mass%C
steel and 1.0mass%C steel. Compared with 0.8mass%C steel, 1.0mass%
C steel revealed less development of plastic flow and crack propagation.
In addition, as shown in Fig. 4, compared with 0.8mass%C steel, 1.0mass
%C steel presented a decrease in spalling formation on the rolling con­
tact surface and a decrease in the crack depth beneath the G.C. portions.
In this section, the reasons for significant improvement of RCF
resistance with increasing carbon content in the repeated dry and water
lubrication conditions are discussed, focusing on the development of
plastic flow and the variation of the crack inclination angles beneath the
rolling contact surface. The results of 0.8mass%C steel with the initial
hardness of 400 HV and 1.0mass%C steel with the initial hardness of 415
HV are considered.

Fig. 11. Relationship between the number of rolling contact cycles and
maximum crack depth beneath the G.C. portion of the rail specimens after
the test.

8
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 12. Relationship between the number of rolling contact cycles and crack inclination angle beneath the G.C. portion of the rail specimens after the test.

4.1. Effect of carbon content on the plastic flow beneath the rolling
σ TS ¼ 9:8 � HV=3 (2)
contact surface
where the unit of σ TS is MPa. The yield strength (σYP) was estimated as
As shown in Figs. 6 and 9, cracks initiate at a portion of the plastic
follows:
flow and propagate along the plastic flow as reported by Sato et al. [3].
Consequently, it is considered that there is a correlation between the σ YP ¼ A � σ TS (3)
development of plastic flow and initiation and propagation of cracks. In
this section, to clarify the effect of carbon content on RCF resistance, the where A is the yield ratio. The shear yield strength (k) beneath the
relationship between the carbon content of rail specimens and growth of rolling contact surface was estimated from the yield strength (σYP) using
plastic flow beneath the rolling contact surface is discussed through the maximum shear strain energy condition (von Mises yield criterion)
stress analysis of the rolling contact surface conducted using the FEM. as follows:
Plastic flow could be significantly affected by tangential traction .pffiffi
exerted on the rolling contact surface and be systematized as a shake­ k ¼ σYP 3 (4)
down map that is calculated and classified from the interrelation of the
The hardness and the estimated values of A, σ YP, k, p0, and p0/k of
tangential traction coefficient, Hertzian maximum contact pressure, and
each specimen are shown in Table 3. Here, the value of A is set at 0.68
shear yield strength of steel [31]. Plastic flow beneath the rolling contact
and 0.90 from the tensile test result of pearlitic steel [32]. The value of
surface would develop under a rolling/sliding contact condition by
pearlitic steel is 0.68 with tensile strength of 1250 MPa, and 0.90 is the
surpassing the shakedown limit. Here, based on the shakedown theory,
value of cold-rolled pearlitic steels with tensile strength from 1400 to
the cause of the difference in the plastic flow of 0.8mass%C steel and
1600 MPa. Herein, in the calculation of σYP in the initial contact con­
1.0mass%C steel is considered.
dition, the value of A is set at only 0.68 for pearlite steel. p0 is the
To begin with, contact stress between the rail specimen and wheel
Hertzian maximum contact pressure. The values of pn calculated by the
specimen under the initial contact condition was calculated by the FEM
FEM were used for the value of p0 in this study.
using Marc (Ver. 2017 1.0). The contact between the rail and wheel was
The value of p0/k in the initial contact condition for 0.8mass%C steel
modeled in the region of a central angle of 30� along the circumferential
was 3.90, and that for 1.0mass%C steel was 3.75. Based on the shake­
direction. Elastic analysis was performed with Young’s modulus of 206
down theory, the p0/k value and the traction coefficient at the contact
GPa and Poisson’s ratio of 0.3 for both rail and wheel materials. In
surface become important factors for evaluating the material response
addition, the conditions of external force were set as same as the test
under the repeated rolling contact condition. With a larger value of p0/k
conditions shown in Table 2. The friction coefficient of the contact
than the prescribed limit, there is a larger possibility to develop plastic
surface was set to 0.2.
flow in the surface layer portion. Furthermore, the shakedown limit
The calculated results are shown in Fig. 15. The rail and wheel
should increase with strain hardening via repeated rolling contact [33].
specimens were principally in contact with each other at the G.C.
The increase of the shakedown limit may tend to suppress the devel­
portion and the head top portion, and the normal contact pressure was
opment of plastic flow if the stress condition does not change.
higher in the G.C. portion than in the head top portion. The maximum
When comparing Table 3 with Fig. 10, it becomes clear that the
contact pressure (pn) at the G.C. portion where spalling occurred in the
plastic flow of both steels develop in the early stage up to the cycle
rail specimens reached 2000 MPa.
number 320 � 103 although the value of p0/k of 1.0mass%C steel is a
Next, shear yield strength (k) of 0.8mass%C steel and 1.0mass%C
slightly smaller than that of 0.8mass%C steel. This fact suggests that at
steel beneath the rolling contact surface was estimated as described
the early stage of the test, the rolling-sliding contact conditions of both
below. Initially, shear yield strength (k) at the initial contact condition
the steels could surpass the shakedown limit.
was calculated by focusing on the initial hardness of 400 HV for 0.8mass
In contrast, with increase in the cycle number up to 420 � 103, the k
%C steel and 415 HV for 1.0mass%C steel. Second, the shear yield
values of both steels increase with strain hardening; however, 0.8mass%
strength during the repeated dry and water lubrication conditions was
C steel with a smaller value of k than that of 1.0mass%C steel has a larger
calculated by considering the change in hardness beneath the rolling
tendency to develop plastic flow in the surface layer during the repeated
contact surface. With respect to the hardness in the repeated dry and
dry and water lubrication conditions. Although it might be difficult to
water lubrication conditions, the hardness of the surface layer portion at
quantitatively discuss the material response of this study based on the
cycle number 420 � 103 was 497 HV for 0.8mass%C steel and 540 HV
shakedown theory at this stage, these results suggest that even an in­
for 1.0mass%C steel, as per the results shown in Fig. 14. Then, the tensile
crease of approximately 9% in the k value with increasing carbon con­
strength (σ TS) was estimated using these values of hardness as follows:
tent of the steels can cause a large difference in the behavior of plastic

9
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 14. Variations in the mean hardness of the surface layer with a thickness of
50 μm at the G.C. portion of the rail specimens and its hardness increase rate
according to the number of rolling contact cycles.

deformation under repeated passage of the rolling load.

4.2. Effect of carbon content on crack propagation beneath the rolling


contact surface

Crack propagation behaviour of surface cracks in lubricated rolling/


sliding contact have been discussed by many researchers on the basis of
the fracture mechanics approach (e.g., Keer et al. [34], Kaneta et al. [35,
36], Bower et al. [37], and Akama et al. [38]). They have shown that the
hydraulic pressure due to fluid penetration into the crack plays an
important role in the stress intensity factors prescribing the intensity of
the stress field near the crack tip. Furthermore, Tyfour et al. [27] and
Kaneta et al. [28] have shown from their experimental studies using a
two-disk-type testing machine that the water lubrication condition after
the dry condition or the repetition of frequent dry and wet condition
strongly affected the crack propagation behaviour.
Fig. 16 shows the schematic model of a surface crack in lubricated
rolling/sliding contact. In this model, Herzian contact pressure p(x) with
surface traction q(x) is assumed to move in the x direction over an elastic
half-space containing a crack with angle of inclination β to the surface; a
and c in this figure are the contact half-width and the crack length,
respectively. At the G.C. portion of the rail specimen, the surface trac­
tion q(x) acts in the opposite direction to load motion.
On the basis of the abovementioned model, each parameter was
established using the results of the present experiments. As shown in
Fig. 15, the contact arc length (2a) in the rolling direction in the G.C.
portion was approximately 3.6 mm.
Here, the crack inclination angle (β) was examined in detail because
the cracks observed in this study do not have an entirely straight shape
as shown in Fig. 16, but form an upward projecting curve. Fig. 17 shows
typical cross-sectional aspect beneath the G.C. portion at cycle number
of 420 � 103 at which the crack propagation behavior tends to differ
depending on the kind of steel under repeated dry and water lubrication
conditions. The lengths of typical cracks were 0.88 and 1.40 mm (880
and 1400 μm) for 0.8mass%C steel and 0.52 and 0.72 mm (520 and 720
μm) for 1.0mass%C steel.
By contrast, the crack inclination angle varies depending on the
method of measurement. When the inclination angle of the line obtained
Fig. 13. Hardness distribution beneath the G.C. portion of the rail specimens: by connecting the crack mouth and the crack tip was measured for the
(a) 320 x 103 cycles, (b) 420 x 103 cycles, (c) 520 x 103 cycles. same cracks of which the length was measured above, the inclination
angle (β T) became 23� and 28� for 0.8mass%C steel and 37� and 36� for
1.0mass%C steel. In contrast, when the line connecting the crack mouth
and the point on the crack at the depth of 200 μm was measured as
defined in Fig. 12, the inclination angle (β 200) became 19� and 20� for

10
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 15. Pressure and stress distributions of the rail specimen in contact with the wheel specimen under the initial condition: (a) Distribution of normal contact
pressure pn on a rail specimen surface, (b) Distribution of σ z along the cross section passing through the line A-A.

Table 3
Comparison between maximum contact pressure calculated by FEM and shear yield
strength estimated based on the initial hardness and hardness beneath the G.C.
portion at cycle number 420 � 103.

0.8mass%C steel and 28� and 32� for 1.0mass%C steel. Clearly, although calculated as 0.48 and 0.78 for 0.8mass%C steel and 0.29 and 0.40 for
the crack inclination angle tends to increase as the measurement depth 1.0mass%C steel, respectively.
increases, the crack inclination angle differs depending on the kind of The influence of water that penetrated into the cracks on the prop­
steel irrespective of the method of measurement, and 0.8mass%C steel agation behavior of the modeled cracks for the present experiments was
showed smaller inclination angle than 1.0mass%C steel. estimated from the analytical results of the stress intensity factor (SIF) in
From cracks whose inclination angles were measured, the value of c/ references based on the fluid entrapment mechanism.
a, which affects the stress concentration at the tip of the cracks, was The crack generation condition and analysis results of the SIF are

11
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 16. Schematic model of a surface crack in lubricated rolling/


sliding contact.

organized as shown in Table 4. Note that the values of the SIF in the table
are maximum when the Hertzian pressure distribution moves over the
surface.
Bower et al. [37] pointed out that when the Hertzian contact pres­
sure works in the vicinity of the mouth of a crack opening spot, both
Mode-I SIF (KI) of the opening mode and Mode-II SIF (KII) of the shear
mode increase, and particularly, the KI increases compared with the KII.
In addition, Akama et al. [38] and Dallago et al. [39] have shown that
the KI increases as the crack inclination angle (β) decreases. In partic­
ular, according to the analytical results by Akama et al., the KI increases
notably when the crack inclination angle (β) reduces below 25� or less.
In contrast, Makino et al. have shown that the Mode-I SIF threshold (KI
th) is smaller than the Mode-II SIF threshold (KII th) when cracks are
generated in wheel materials exhibiting a pearlitic structure [40].
Therefore, it is surmised that an increasing KI at the tip of the cracks
facilitates crack propagation.
The crack inclination angle is considered to be affected by the wear
Fig. 17. Typical cross-sectional aspects beneath the G.C. portion of rail speci­
of the rolling contact surface. As shown in Fig. 8, 0.8mass%C steel has a
mens after 420 � 103 cycles: (a) 0.8mass%C 400 HV, (b) 1.0mass%C 415 HV.
large amount of wear compared with 1.0mass%C steel. If the wear does
not occur, the crack inclination angle of 0.8mass%C steel would
such that plastic flow tended to develop. Meanwhile, in 1.0mass%C
decrease furthermore which could result in the increase of value of the
steel, the hardness beneath the rolling contact surface was higher due to
KI at the tip of the crack.
the greater extent of HIR, resulting in the increase in the shear yield
With these results, it is presumed that 0.8mass%C steel with a lower
stress (k). In this respect, in 1.0mass%C steel, the surface layer portion
carbon content and smaller crack inclination angle was under the status
beneath the rolling contact surface was less likely to exceed the shake­
in which crack propagation tends to be facilitated by significantly
down limit, thereby inhibiting the development of plastic flow.
increasing the KI at the tip of the cracks due to water penetration into the
With the development of plastic flow, the cracks propagated during
cracks. Conversely, it is surmised that 1.0mass%C steel with a higher
water lubrication or subsequent repeated dry and water lubrication
carbon content and larger crack inclination angle was under the status in
conditions should incline along the plastic flow. Due to this situation, the
which crack propagation is unlikely to be promoted compared with
crack inclination angle is likely to decrease in 0.8mass%C steel with
0.8mass%C steel.
large development of plastic flow and unlikely to decrease in 1.0mass%C
steel. As a result, when water penetrates into the cracks during wet
4.3. Possible mechanism of spalling in pearlitic steel having different
carbon content
Table 4
Based on the abovementioned considerations, a schematic illustra­ Estimated c/a and β in this study and their effects on KI and KII obtained by
tion of structural variation and behavior of crack propagation, as well as Bower [37] and Akama [38].
occurrence of spalling beneath the G.C. portion of the rail specimen, is Results c/a β (� ) K I/ K II/
shown in Fig. 18. β 200, β T (p0√a) (p0√a)
(� )
At the end of water lubrication, plastic flow and micro-cracks were
generated in both 0.8mass%C steel and 1.0mass%C steel, where no Present 0.8mass% Crack- 0.48 19, 28 – –
significant difference was identified in the plastic flow behavior and C 1
Crack- 0.78 20, 23
crack growth.
– –
2
However, subsequently, during dry running in the repeated dry and 1.0mass% Crack- 0.29 32, 37 – –
water lubrication conditions, the friction coefficient increased on the C 1
rolling contact surface in the G.C. portion such that plastic flow beneath Crack- 0.40 28, 36 – –
2
the rolling contact surface developed. Compared with 1.0mass%C steel,
Bower 0.50 25 0.16 0.14
0.8mass%C steel with a lower HIR had low hardness beneath the rolling [37]
contact surface and low shear yield strength (k). Here, the large amount Akama 0.50 50 0.11 0.18
of wear for 0.8mass%C steel could also encourage the lower HIR. [38]
Therefore, in 0.8mass%C steel, the surface layer portion beneath the 25 0.18 0.12
12.5 0.30 0.07
rolling contact surface was most likely to exceed the shakedown limit

12
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

Fig. 18. Schematic illustration of structural variation, behaviour of crack propagation, and occurrence of spalling beneath the G.C. portion of the rail specimens.

running, the magnitude of KI of 0.8mass%C steel becomes larger than (1) The number of spalls on the rolling contact surface and the depth
that of 1.0mass%C steel, then the tendency of mode I crack propagation of the cracks generated beneath the rolling contact surface
in 0.8mass%C steel surpasses that in 1.0mass%C steel. This results in the decrease with increasing the carbon content of pearlitic steel even
difference in crack depth between them despite a larger amount of wear when the initial hardness is almost the same level.
in 0.8mass%C steel. (2) The increase in carbon content of the pearlitic steel increases the
The crack growth direction by tensile mode can be predicted using hardness beneath the rolling contact surface, resulting in the
Kσθ, which is the SIF that affects the intensity of the field of normal suppression of the development of plastic flow.
stress near the crack tip. Kσθ is defined as follows [41]: (3) The crack inclination angle is correlated with the development of
�θ � � �θ � 3 � plastic flow. That is, the pearlitic steel with high carbon content
KσðθÞ ¼ cos KIcos2 KIIsinðθÞ (5) has a large crack inclination angle from the rolling contact sur­
2 2 2
face than that with low carbon content.
(4) As the decrease in the crack inclination angle from the rolling
where θ is the angle from the line of the crack as shown in Fig. 16.
contact surface can increase the stress concentration at the tip of
Dallago et al. [39] have shown from their analytical study that Kσθ in­
the cracks due to water penetration into the cracks, pearlitic steel
creases due to water penetration into the cracks and that the value of
with high carbon content has a larger tendency to suppress crack
Kσθ is maximum in the direction toward the upper surface side from the
propagation compared with that with low carbon content. As a
tip of the cracks, i.e., the rolling contact surface. In addition, Ancellotti
result, the RCF resistance is improved.
et al. [42] and Makino et al. [40] have indicated that in cases where the
length of cracks is long as c/a ¼ 4.0, even under low water pressure
generated in inside the cracks, the value of Kσθ becomes maximum in Declaration of competing interest
the direction of the rolling contact surface. As the value of KII th is
smaller than the value of KI th [40], cracks tend to branch toward the The authors declare that they have no known competing financial
rolling contact surface. Accordingly, compared with 1.0mass%C steel, interests or personal relationships that could have appeared to influence
0.8mass%C steel in which cracks deeply propagate and the length of the work reported in this paper.
crack increases tends to allow the cracks to branch toward the direction
of the rolling contact surface. Then, such branched cracks coalesce with Acknowledgment
the surrounding cracks, facilitating spalling on the rolling contact
surface. We wish to express our sincere appreciation toward Dr. Yamamoto
According to the mechanism described above, it is surmised that RCF (NIPPON STEEL TECHNOLOGY Co., Ltd) who cooperated with us for
resistance of 1.0mass%C steel is significantly improved compared with conducting the contact stress analysis of the rail and wheel specimens in
that of 0.8mass%C steel. this study.

5. Conclusion References

The RCF resistance of rails was evaluated by conducting tests in [1] L. Wessels, S. Oswald, D. Welsby, P. Mutton, Managing the transition from rail
which the rolling contact between the rail and wheel was reproduced wear to rolling contact fatigue in a heavy haul environment, in: Proceedings of the
International Heavy Haul Conference, Perth, 2015, pp. 1045–1054.
using pearlitic steel with varying carbon content and hardness values. [2] H. Ghonem, J. Kalousek, D.H. Stone, E.E. Laufer, Aspects of plastic deformation
The following conclusions can be drawn: and fatigue damage in pearlitic rail stee, in: Proceedings of the 2th International
Heavy Haul Conference, Colorado Springs, 1982, pp. 339–349.

13
M. Ueda and K. Matsuda Wear 444-445 (2020) 203120

[3] M. Sato, P.M. Anderson, D.A. Rigney, Rolling-sliding behavior of rail steels, Wear [23] D. Li, C. Pinney, S. Kalay, D. Davis, Recent advances in rail life extension in north
162–164 (1993) 159–172. American heavy haul railways, in: Proceedings of the International Heavy Haul
[4] H. Muster, H. Schmedders, K. Wick, H. Pradier, Rail rolling contact fatigue. The Conference, Calgary, 2011.
performance of naturally hard and head-hardened rails in track, Wear 191 (1996) [24] A.R. Marder, B.L. Bramfitt, The effect of morphology on the strength of pearlite,
54–64. Metall. Trans. A 7A (1976) 365–372.
[5] P. Clayton, Tribological aspects of wheel-rail contact: a review of recent [25] B.E. O’Donnelly, R.L. Reuben, T.N. Baker, Quantitative assessment of
experimental research, Wear 191 (1996) 170–183. strengthening parameters in ferrite-pearlite steels from microstructural
[6] E. Magel, P. Mutton, A. Ekberg, A. Kapoor, Rolling contact fatigue wear and broken measurements, Met. Technol. 11 (1984) 45–51.
rail derailments, Wear 366–367 (2016) 249–257. [26] M. Ueda, K. Uchino, A. Kobayashi, Effects of carbon content on wear property in
[7] J. Kristan, L. Allen, J. LoPresti, Evaluation of advance rail steels and improved pearlitic steels, Wear 253 (2002) 107–113.
welding techniques under 35.7-tonne (39 ton) axle loads at the facility for [27] H.L. Heathcote, The ball bearing: in the making, under test and on service, Proc.
accelerated service testing (FAST), transportation technology center, USA, in: Inst. Automob. Eng. 15 (1920) 569–702.
Proceedings of the 8th International Heavy Haul Conference, Rio de Janeiro, 2005, [28] W.R. Tyfour, J.H. Beynon, A. Kapoor, Deterioration of rolling contact fatigue life of
pp. 709–717. pearlitic rail steel due to dry-wet tolling-sliding line contact, Wear 197 (1996)
[8] J. Duvel, P. Mutton, E. Alvarez, J. Mcleod, Rail requirements for 40 tonne axle 255–265.
loads, in: Proceedings of the 8th International Heavy Haul Conference, Rio de [29] M. Kaneta, K. Matsuda, K. Murakami, H. Nishikawa, A possible mechanism for rail
Janeiro, 2005, pp. 719–729. dark spot defects, J. Tribol. 120 (1998) 304–309.
[9] Van K. Dang, M.H. Maitournam, On some recent trends in modelling of contact [30] D.T. Eadie, D. Elvidge, K. Oldknow, R. Stock, P. Pointner, J. Kalousek, P. Klauser,
fatigue and wear in rail, Wear 253 (2002) 219–227. The effects of top of rail friction modifier on wear and rolling contact fatigue: full-
[10] H. Desimone, On the application of Dang Van criterion to rolling contact fatigue, scale rail–wheel test rig evaluation, analysis and modelling, Wear 265 (2008)
Wear 260 (2006) 567–572. 1222–1230.
[11] D.F.C. Peixoto, Application of the Dang van fatigue criterion to the rail/wheel [31] K.L. Johnson, Contact Mechanics, Cambridge University Press, 1985.
contact problem, Mater. Sci. Forum 636–637 (2010) 1178–1185. [32] Y. Liua, Effects of microstructure and crystallography on mechanical properties of
[12] R.P. Reiff, Rail/wheel Lubrication Studies at FAST, Lubrication Engineering, 1986, cold-rolled SAE1078 pearlitic steel, Mater. Sci. Eng. A 709 (2018) 115–124.
pp. 340–349. June. [33] C.P. Jones, W.R. Tyfour, J.H. Beynon, A. Kapoor, The effect of strain hardening on
[13] P.J. Mutton, Aspects of rail lubrication in heavy haul railroads, Mech. Eng. Trans. shakedown limits of a pearlitic rail steel, Proc. Inst. Mech. Eng. 211 (1997)
13 (3) (1988) 157–164. 131–140.
[14] T. Judge, Rail Lubrication Providing Potential, Railway Track & Structure, 2009, [34] L.M. Keer, M.D. Bryant, A pitting model for rolling contact fatigue, ASME J. Lubr.
pp. 27–32. February. Technol. 105 (1983) 198–205.
[15] H.C. Eden, J.E. Garnham, C.L. Davis, Influential microstructural changes on rolling [35] M. Kaneta, Y. Murakami, Effects of oil hydraulic pressure on surface crack growth
contact fatigue crack initiation in pearlitic rail steels, Mater. Sci. Technol. 21 (6) in rolling/sliding contact, Tribol. Int. 20 (No.4) (1987) 210–217.
(2005) 623–629. [36] M. Kaneta, Y. Murakami, Propagation of semi-elliptical surface cracks in lubricated
[16] J.E. Garnham, C.L. Davis, The role of deformed rail microstructure on rolling rolling/sliding elliptical contacts, ASME J. Tribol. 113 (1991) 270–275.
contact fatigue initiation, Wear 265 (2008) 1363–1372. [37] A.F. Bower, The influence of crack face friction and trapped fluid on surface
[17] J.H. Beynon, J.E. Garnham, K.J. Sawley, Rolling contact fatigue of three pearlitic initiated rolling contact fatigue cracks, ASME J. Tribol. 110 (1988) 704–711.
rail steels, Wear 192 (1996) 94–111. [38] M. Akama, T. Mori, Boundary Element Analysis of K-Value for Rolling Contact
[18] R. Stock, R. Pippan, RCF and wear in theory and practices-The influence of rail Fatigue Cracks, No.20070012, Transaction of JSCES, 2007 [in Japanese].
grade on wear and RCF, Wear 314 (2011) 125–133. [39] M. Dallago, M. Benedetti, S. Ancellotti, V. Fontanari, The role of lubricating fluid
[19] R. Stock, R. Pippan, Rail grade dependent damage behavior -Characteristics and pressurization and entrapment on the path of inclined edge cracks originated under
damage formation hypothesis, Wear 314 (2014) 44–50. rolling–sliding contact fatigue: numerical analyses vs. experimental evidences, Int.
[20] G. Girsch, N. Frank, P. Pointner, New rail grade – a technical performance J. Fatigue 92 (2016) 517–530.
overview, in: Proceedings of the 8th International Heavy Haul Conference, Rio de [40] T. Makino, T. Kato, K. Hirakawa, The effect of slip ratio on the rolling contact
Janeiro, 2005, pp. 731–738. fatigue property of railway wheel, Int. J. Fatigue 36 (2012) 68–79.
[21] K. Iwano, M. Ueda, K. Karimine, T. Yamamoto, Recent development of rails in [41] F. Erdogan, G.C. Sih, On the crack extension in plates under plane loading and
nippon steel, in: Proceedings of 7th International Conference of Contact Mechanics transverse shear, J. Basic Eng. 85 (1963) 519–527.
and Wear of Rail/wheel Systems (CM 2006), 2006, pp. 287–293. Brisbane. [42] S. Ancellotti, M. Benedetti, M. Dallago, V. Fontanari, The role of the second body
[22] M. Ueda, K. Iwano, T. Yamamoto, Rail performance and recent developments of on the pressurization and entrapment of oil in cracks produced under lubricated
rail, in: Proceedings of the International Heavy Haul Association Conference, rolling-sliding contact fatigue, Theor. Appl. Fract. Mech. 91 (2017) 3–16.
Calgary, 2011.

14

You might also like