Anda di halaman 1dari 12

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/315812760

Dry and lubricated wear of rail steel under


rolling contact fatigue-Wear mechanisms and
crack growth

Article · May 2017


DOI: 10.1016/j.wear.2017.03.025

CITATIONS READS

0 17

3 authors, including:

Santiago Maya Johnson Juan Felipe Santa


Universidade Federal de São Carlos Instituto Tecnológico Metropolitano
8 PUBLICATIONS 7 CITATIONS 21 PUBLICATIONS 153 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Santiago Maya Johnson on 07 April 2017.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Wear 380-381 (2017) 240–250

Contents lists available at ScienceDirect

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

Dry and lubricated wear of rail steel under rolling contact fatigue -
Wear mechanisms and crack growth
Santiago Maya-Johnson a,b,n, Juan Felipe Santa a,c, Alejandro Toro a
a
Tribology and Surfaces Group, National University of Colombia, ZIP 050041 Medellín, Colombia
b
Brazilian Nanotechnology National Laboratory (LNNano), ZIP 13083-970 Campinas, Brazil
c
Grupo de Investigación Materiales Avanzados y Energía – MATyER, Instituto Tecnológico Metropolitano, Medellín, Colombia

art ic l e i nf o a b s t r a c t

Article history: In this work, the dry and lubricated wear behavior of rail steels (R370CrHT) under Rolling Contact Fatigue
Received 16 October 2016 (RCF) conditions was studied in laboratory. A twin-disc rolling-sliding machine was used to reproduce
Received in revised form contact conditions under high contact pressure (1.1 GPa) and high creepage (5%) with the aim of ac-
11 February 2017
celerating RCF failure. Two different types of tests were performed, namely dry and lubricated with a
Accepted 28 March 2017
friction modifier, to simulate the mechanisms that cause RCF. The results showed an increase of the wear
Available online 31 March 2017
rate in the tests where a friction modifier was added after an initial stage of 4,000 dry cycles. The damage
Keywords: of the surface proceeded by cracks formation during the dry stage followed by accelerated crack growth
Rolling contact fatigue and flaking due to the effects induced by the friction modifier entering into the original cracks. After
Rail-wheel tribology
several thousand cycles, the flakes were detached from the surface causing high wear rates. The cracks
Friction modifiers
morphology was observed under SEM. The cracks depth and angle with respect to the contact surface
Twin disc testing
Fluid pressure were reported and some correlations were made with the mass loss results.
& 2017 Elsevier B.V. All rights reserved.

1. Introduction A research in Australia found that the total annual maintenance


cost of a 12 Million Gross Tons (MGT) railway system can be as
The wear on rail systems is a serious problem for the compe- high as USD $ 34 per meter of rail. The costs and the wear rates are
titiveness of transport industry. A global trend to address wear influenced by the curvature radius of the railroad, the planning
issues is the use of friction modifiers (FMs) at the interface be- intervals of rail grinding (Rail grinding refers to a controlled wear
tween wheel and rail. Bower (1988) [1], Kaneta and Murakami operation used to eliminate cracks before they reach a critical
(1985) [2–4], and others have studied rolling contact fatigue (RCF) length at which its growth rate sharply increases) and the use of a
to gain understanding of the behavior of the FM when entering the suitable FM at the interface, among others. The total annual cost/
cracks and to evaluate its influence on wear and cracks growth meter of the rail maintenance in a non-lubricated 12 MGT trans-
[5–9]. Bower [1] proposed three mechanisms of crack growth as- port systems is USD $ 54, whereas with lubrication savings could
sisted by fluids based on the experimental results of S. Way amount around 60% of the total value [10].
(1935): (i) the cracks only propagate if a fluid lubricant is applied A survey released by the Engineering Research Programme in 2003
to the surfaces in contact, (ii) the cracks always propagate in the [11], where the information gathered from 35 rail networks was
direction of motion of the load over the surface, and (iii) if there is analyzed, concluded that effective lubrication can reduce wear on
some relative sliding between the two contacting surfaces, cracks wheel and rail and the noise levels. The rails life can be increased by a
only propagate in the driven surface. Although it is acknowledged factor of two, the wheel life by a factor of five and in some cases
that a FM can minimize friction between the crack faces and depending on the radius of curvature and grinding intervals by a
promote crack growth through the Mode II (shear), it is also well factor of four. However, today many of the steps and principles that
known these days that no fluid is required to grow a crack; re- govern the phenomenon of RCF in the presence of a FM are not en-
garding (iii), it has been found that the cracks grow in both wheel tirely known. Several studies carried all over the world have helped
and rail. understand some of the principles that govern the phenomenon of
RCF and other problems related to the tribology of wheel/rail contact
n
such as the type of predominant defects, nondestructive techniques
Corresponding author at: Tribology and Surfaces Group, National University of
Colombia, ZIP 050041 Medellín, Colombia.
for inspecting rails, among others [12–18].
E-mail addresses: smayaj@gmail.com (S. Maya-Johnson), In this work, the dry and lubricated wear behavior of rail steels
jfsanta@gmail.com (J. Felipe Santa), aotoro@unal.edu.co (A. Toro). under RCF conditions was studied in laboratory. A twin-disc

http://dx.doi.org/10.1016/j.wear.2017.03.025
0043-1648/& 2017 Elsevier B.V. All rights reserved.
S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250 241

Table 1
Chemical composition of rail and wheel (wt%).

C Si Mn P S Cr Al V Cu Ti Ni Mo

R370CrHT 0.762 0.394 1.062 0.011 0.015 0.506 0.000 0.002 0.038 0.002 0.058 0.019
ER8 0.542 0.253 0.734 0.011 0.006 0.141 0.027 0.006 0.165 0.002 0.120 0.048

rolling-sliding machine was used to impose controlled contact


Table 2
pressure and creepage conditions that led the tribological pair to a Mechanical properties of the rail.
severe wear regime. The study was done to evaluate the effect of
the application of a FM to the interface when cracks are already Tensile Yield Strength Elongation (%) Hardness HV
formed at the surface of the rail. Strength (MPa) (HB)
(MPa)

R370CrHT 1373 767 10 387 7 17


2. Materials and methods (366.4)

Rail specimens were extracted from R370CrHT rail sections,


manufactured by the company Voestalpine Schienen GMBH-Austria, Fletcher and Beynon in a R260 rail steel [23], the number of dry cycles
and wheel specimens were extracted from a railway wheel grade must be carefully controlled since high tangential forces cause the
ER8. The chemical composition measured by optical emission cracks to grow rapidly and severe flaking may occur after 5000 cycles,
spectroscopy and relevant mechanical properties of the rail and so the nucleated cracks are lost. The number of dry cycles was set
wheel, in line with European standards EN 13674-1:2011 [19] and equal for all samples, since Tyfour et al. [8] demonstrated that this has
EN 13262:2004 [20], are shown in Table 1 and Table 2 respectively. a notable effect on the final fatigue life in dry-wet contact.
The hardness of the wheel was 291 717 HV31.25kgf. Accordingly, in this study, 4,000 dry cycles of pre-fatigue were
The samples were polished for metallographic analysis by using used based on the results of J.F. Santa [24]. Without stopping the
emery papers and (6 to 0.25 μm) diamond paste. After fine pol- test, after 4,000 dry cycles a commercial FM (Sintono Terra HLK,
ishing, the samples were etched with Picral (100 ml ethanol þ 4 g based on synthetic oils) was added to the surface of the wheel
specimen by using a brush. The amount of FM added was weighed
picric acid). A JEOL 5910LV SEM was used to evaluate the micro-
structure and the worn surfaces using Secondary Electrons (SE). to ensure that 0.02 to 0.03 g would be applied to the wheels sur-
For the longitudinal sections and EDS mapping a FEI Inspect F50 face. The application was repeated every 400 cycles. Once the FM
is applied, the coefficient of traction (COT) drops sharply to reach a
was used, located at the Brazilian Nanotechnology National La-
boratory (LNNano). To differentiate the inclusions from the cracks stable value after a few tens of cycles.
Three different numbers of cycles were selected to evaluate the
Backscattered Electrons (BSE) images were taken.
effect of the FM on RCF: 9,000, 14,000, and 24,000. For comparison
The wear tests were carried out in a Twin-Disk testing machine
reasons, dry tests with equal number of total cycles were also
installed at the Tribology and Surfaces Laboratory in the National
performed. For each condition two replicas were prepared, and
University of Colombia, Medellín. Constructive details about the
before each test the samples were washed and weighed to mea-
machine can be found elsewhere [21]. The Twin-Disk machine is
sure mass losses. The tests were performed using separate sets of
used to simulate either dry or wet contact in wheel/rail tribosys-
samples for each number of cycles to be tested. This is because
tems [22]. For each condition, two replicas were performed on
stopping the tests would induce a new running-in period during
different (new) samples to evaluate the repeatability. In this study,
the re-start due to misalignment, changes in real contact area and
the lubricated condition was provided by the addition of a FM to
roughness accommodation. W.R. Tyfour et al. show how the COT
the contact interface. In the test rig, two rolling discs are loaded
never reaches a stable value if the tests are stopped and re-started
upon each other and the relative speed, contact pressure and slip
with the same pair of samples [25].
percentage (creepage) are precisely controlled. Fig. 1 shows a
At the beginning of every test the applied load and contact area
schematic of the testing device.
were verified by carefully watching the contact path in the sam-
Cylindrical specimens were extracted from actual wheels and
ples with the applied load. After turning the motors on, the friction
rails by cutting and lathe turning. The rail specimens were ex-
torque was corrected to zero, the load was applied and the test
tracted from the rails head (Fig. 2a). Both rail and wheel specimens
started. Once the test ended, the samples were washed with de-
have the same dimensions, with a diameter of 47 mm and a rolling
greaser, submerged in acetone and cleaned in ultrasonic bath to
width of 10 mm. The experiments were designed to be aggressive
remove the residues of FM. Finally the samples were weighed to
enough to cause RCF under severe wear regime, so a slip percen-
measure the mass losses and to obtain the wear rates.
tage of 5% and a contact pressure of 1.1 GPa were chosen. The
wheel specimens were rotating at 401 71 RPM, and the rails
specimens at 380 71 RPM. As the samples rotate at different RPM
3. Results and discussion
they are subjected to different fatigue cycles; for the diagrams the
cycles were calculated according to the wheel disc revolutions (i.e.
The nomenclature used hereinafter has the first letter related to
400 RPM). The slip percentage was defined according to Fletcher
the interfacial condition (H for HLK and D for dry), followed by the
and Beynon as [23]:
slip percentage and the number of wet cycles. As an example, a
⎛R N −R N ⎞ test run with the addition of HLK with 1.1 GPa, 5% creepage and
Slip( %) = 200⎜ W W R R

⎝ RWNW + RRNR ⎠ (1) 10,000 wet cycles, is named “H5% 10 K”.

where N are the revolutions, R is the radius of the disk specimens, 3.1. Effect of the use of a FM on the RCF
and W and R indicate wheel and rail respectively.
In order to accelerate the cracks growth, a dry pre-fatigue period In Fig. 3 the COT reaches a maximum of 0.63 at the beginning of
was implemented before applying the FM. As it was shown by the test and then it decreases to values below 0.60. The running
242 S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250

Fig. 1. Schematic of the Twin-Disk testing machine.

the surface and to reach the dry friction coefficient. For most of the
tests, the COT is close to the dry condition after one application, so
for the quantity of FM applied, the retentivity is around 400 cycles.
After five applications (1600 cycles) the COT was very stable.
In order to prevent influencing the COT by the retentivity, all
data used to calculate the COT were averaged starting from 6,600
cycles (Table 3). In both conditions, lubricated with FM and dry, a
low standard deviation was observed in the COT confirming that
the testing machine is reliable. For the tests with FM the average
COT is E0.07 while in dry condition is E 0.49. Field measure-
ments in a commercial railway with similar tests conditions
showed COT between 0.13 and 0.15 [27]. The variation between
laboratory tests and field measurements of friction have been
studied for years, some assumptions have been made but the
Fig. 2. a) Extraction of the rails samples for tribological tests, b) cutting scheme for reasons are still unknown [28,29]. To ensure that the HLK is able to
subsurface analysis.
guarantee the minimum COT to maintain the operation conditions
in a railway system (0.20), new tests simulating the top of rail/
wheel tread interface are necessary. The results with this FM are in
accordance with those from Gallardo-Hernandez, et al. [30], and
other tests with R260 at 1.5 GPa, dry and using a grease as FM
[31,32].

3.2. Mass losses

Fig. 4a shows the combined (rail þ wheel) mass losses after


tribological tests with 1.1 GPa and 5% creepage. The rails mass loss
after the application of the FM tended to be slightly higher than
that observed during the dry tests due to flaking and cracks
growth assisted by the fluid. During dry tests, the mass loss of the
wheel caused by abrasion was higher so the combined mass losses
during the dry tests become higher than in the case when a FM
was used. Up to 24,000 cycles, the mass loss in dry conditions
caused by abrasion is higher than the mass loss by fatigue and
flaking, even though the HLK has been added to a previously
cracked surface. The difference becomes smaller with increasing
fatigue cycles as more cracks grow and flakes are detached from
Fig. 3. Comparison of traction curves for wet and dry tests after 9000 fatigue
cycles. the surface. When the cracks length reached during the dry cycles
is greater, the mass losses are higher. S.R. Lewis et al. also reported

track appearance indicates that the mechanism by which the rail


Table 3
disc loses material prior to the steady state stage is oxidative wear, Coefficients of traction.
where the disc track was observed to be covered by a thin, brown
oxide layer [8,26]. When the FM is applied the COT suddenly re- Test condition Total fatigue cycles Coefficient of Traction (COT)

duces from E 0.52 to E0.10 and then increases again once the FM
HLK 9,000 0.077 7 0.007 0.069 7 0.008
is removed from the contact zone. Centrifugal forces during the 14,000 0.068 7 0.004
test and the squeezing effect caused by contact pressure are re- 24,000 0.061 7 0.001
sponsible for removing the FM from the contact patch. This be- Dry 9,000 0.498 7 0.008 0.493 7 0.012
14,000 0.489 7 0.024
havior can be characterized by the retentivity of the FM, which is
24,000 0.493 7 0.003
defined as the number of cycles required to remove the FM from
S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250 243

Fig. 4. (a) Wheel and rail mass loss, and (b) wear rate of rail samples for wet and dry tests.

Fig. 5. Images of the cleaned discs after lubricated with FM, and dry tests.
244 S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250

higher wear rates with lubrication starvation in comparison to dry amounts, when compared with the samples with FM. The surface
tests [32]. is very smooth and shiny. On the contrary, wheel samples have
Fig. 4b shows the rails mass loss per cycle of fatigue (i.e. the surfaces with significant damage due to abrasion. The largest
wear rate). The wear rates were calculated as the total mass loss abrasion marks and the higher mass losses were expected in the
divided by the total fatigue cycles (dry and wet). It is interesting wheel specimen, since it has lower hardness than the rail speci-
that during dry tests the wear rate remains almost constant men (between 245–275 HB).
(E 6 mg/cycle), which is in agreement to the results of W.R. Tyfour The formation of cracks in rail samples during the tests with
et al., who showed that wear rate stabilized after a certain number FM is caused by several phenomena: (1) the fluid assisted me-
of fatigue cycles. For their tests at 1.5 GPa and 1% creepage the chanism [1,36], (2) the state of stress in the rolling sliding contact
wear rate stabilized after 15,000 rolling cycles [25]. Comparing the [37,38], (3) the reduction of friction within the cracks faces [1,36]
wear rate with the crack angle (Fig. 12b) on the dry tests, between and (4) squeeze film lubrication theory [39]. When a disc slides on
14,000 and 24,000 cycles it seems to be a stable state with no another disc it appears a tangential force that varies with the slip.
change on the crack angle (10 degrees) or the wear rate. Never- The nature of the behavior of the traction-creepage curve was
theless, to clearly identify where the wear rate begins to stabilize described by U. Olofsson and R. Lewis [17,18]. These tangential
for these conditions, more tests with different cycles are needed. forces cause compressive and tensile circumferential stresses in
For the tests with FM the wear rate of the rails is only constant and around the contact zone, and the stresses are oppositely lo-
between 9,000 and 14,000 cycles while after 24,000 cycles the cated in the wheel and rail (Fig. 6). Prior to the contact zone the
wear rate is almost twice as high. If the rails total mass losses at compressive stresses close the tips of the cracks in the wheel
14,000 cycles is subtracted from the mass losses at 24,000 cycles avoiding the fluids to enter in the cracks. On the other hand, the
and the result is divided by the difference in cycles (10,000 cycles), tensile stresses open the cracks tips in the rail and facilitate the
a wear rate almost four times higher than in the early stages is entrance of fluids (Fig. 6b); right after the contact zone the stresses
found (20.185 mg/cycle). The mechanism responsible for the higher are reversed and the crack in the rail closes pressurizing the fluid
wear rates is delamination of big cracks causing spalling (RCF). inside (Fig. 6c).
In Fig. 7 and Fig. 8 the images (a, c, e) correspond to a tangential
3.3. Wear mechanisms view of the surface and (b, d, f) to the normal view; the zoom
inside the images is 200X and 1000X respectively. Fig. 7 shows the
Fig. 5 shows the worn surfaces of rail and wheel samples after evolution of the worn surfaces lubricated with HLK. From 9,000 to
the tribological tests. For the samples with FM the main wear 14,000 cycles, it is observed an increase in the number of cracks by
mechanism on the rail is RCF and small cracks are observed on the RCF on the rail samples. From 14,000 to 24,000 a detachment of
surface. The cracks (head checks) are very similar to those re- the deformed material occurs by the coalescence of the cracks
ported on the literature from field inspections [33,34] and in full- below the surface. This detachment explains the increase in the
scale test rig [35]. The cracks can be visible to the naked eye after mass loss per cycle of the rail samples between 14,000 and 24,000
10,000 cycles of fatigue with FM. After 20,000 cycles with FM, head wet cycles (Fig. 4b).
checks were observed and spalling was also present on the surface. In the dry tests less cracks were observed, Fig. 8. The cracks on
It is worth noticing that no cracks were observed on the wheel the dry tests (Fig. 8c and e) appear shallower than in the wet tests
specimen and only a few abrasion marks were found. (Fig. 7c and e). This is due to a higher tangential force on the dry
In the dry tests, the rail samples appear to have no cracks or tests producing more deformation nearer the surface. The fact that
any other marks related to RCF when observed at the naked eye. there are less cracks on the dry samples than on the wet samples is
Further observations with SEM showed shallow cracks in lesser due two factors: 1) adding a fluid to a system with pre-existing

Fig. 6. Representation of: (a) the state of stress in the two-discs in rolling-sliding condition, (b) closed crack in wheel and opened crack in rail prior to contact zone, and,
(c) opened crack in wheel and closed crack with entrapped fluid in rail. Characteristic flake and crack orientation are show (not to scale) for each disc. Adapted from Garnham
and Beynon (1992) [38].
S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250 245

Fig. 7. SE images of the worn surfaces of the rail samples after tribological test with 1.1 GPa and 5% creepage. 4,000 dry cycles followed by: a) and b) 5,000, c) and d) 10,000,
and e) and f) 20,000 wet cycles with HLK.

cracks promotes the crack growth via the fluid pressurization small cracks that are difficult to detect or are detached during
phenomenon, and 2) the shallower cracks on the dry tests are polishing. Thereby, it is a fact that the samples with FM have more
easily detached, which prevents its subsequent growth, so not cracks and are homogeneously distributed.
many cracks are observed. Deformation of the inclusions in the direction of strain was ob-
Comparing the samples tested under dry and lubricated with served in all samples; the inclusions appear as gray patches and are
FM conditions after 9,000 cycles (Fig. 7a and Fig. 8a), the differ- easily identifiable on the full resolution images (available online).
ences between the tangential views are minor. Thus, applying a Further EDS analysis revealed that the inclusions are mainly sulfides
FM after 5,000 fatigue cycles does not have much influence on the of the MnS type (Fig. 9). When the material becomes more strained
crack growth. Nevertheless, as will be seen on the longitudinal the inclusions deform and align with the cracks and assist their
cuts, the cracks depth and angle remains higher on samples with growth. This behavior was previously reported with pro-eutectoid
FM. ferrite and ductile inclusions in non-fully pearlitic rail steels [40,41].
Previous studies have shown that the presence of inclusions
3.4. Crack growth and their orientation affect fatigue crack growth [42–44]. Fully
pearlitic steels with fewer inclusions have shown lower fatigue
Fig. 10 and Fig. 11 show an overview of the longitudinal section crack growth rate, even without a strained microstructure [44].
of the rail samples, the scan angle was around 35 degrees. The When the microstructure is strained a higher content of ductile
number of cracks observed on the figures is qualitative, since the inclusions, deformed and aligned in the direction of the crack tip,
cracks distribution in samples tested with FM is uniform over the can act as stress concentrators increasing the crack growth.
entire circumference. On the other hand, only a few cracks were With more fatigue cycles an increase of the crack depth is ob-
observed in the dry samples and the selected scanning area had served for both conditions, dry and with FM. On the samples with
the highest density of cracks. Also, the samples with FM had very FM an increase of the crack angle is also observed, but for the dry
246 S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250

Fig. 8. SE images of the worn surfaces of the rail samples after tribological test with 1.1 GPa and 5% creepage. a) and b) 9,000, c) and d) 14,000, and e) and f) 24,000 dry
cycles.

samples between 14,000 and 24,000 fatigue cycles seems to sta- deformation when grease is used.
bilize around 10° (Fig 12b). Fletcher and Beynon carried out tests in R220 with 1% creepage
N. Larijani et al. evaluated the effect of anisotropy in mechanical and 1.5 GPa. They performed three types of tests: unlubricated dry
properties of the surface layer of pearlitic steels induced by large cycles, 250–500 dry cycles plus lubrication using molybdenum
deformations, using a numerical model [45]. They showed that for disulphide suspended in oil, and 250–500 dry cycles plus lu-
isotropy or low degrees of anisotropy, the crack grows towards the brication using water; obtaining crack depths and crack angles of
surface while for high degree of anisotropy the crack grows 11 μm and 10°, 300–360 μm and 10–20°, and 340–550 μm and 20–
downwards. On the highest levels of anisotropy the crack grows on 30° respectively [46]. As in our results, an initial unlubricated
the same angle up to 400–600 μm before an abrupt change to a period followed by lubrication produce large cracks with a higher
vertical direction, this coincides with the results observed for dry angle at the crack mouth. Different factors govern the crack
tests which have higher anisotropy and they are in early stages of growth such as reduction of cracks face friction, pressurization
growth. In all our tests 4,000 dry cycles of pre-fatigue were used, effects, contact forces, squeeze film theory and anisotropy of the
so all start with the same anisotropy, but the wet samples (with properties due to deformation, among others. The water has a
low anisotropy, low strain) have higher crack angle, which does higher COT than others FMs, such as molybdenum disulphide or
not coincide with the results of N. Larijani et al. This shows that not HLK, so the tangential forces and the anisotropy will be higher
only the anisotropy is important to predict crack propagation but when water is used as a FM, but less than unlubricated. In that
also the magnitude of the forces and the pressurization effects case it seems that pressurization effects and the anisotropy play a
have to be considered. Hardwick et al. measured subsurface de- more important role on the crack growth than the crack face
formation on R260 at three slip levels (0.5, 5 and 20%), under dry, friction and the contact forces, since cracks grow faster with water
water and grease lubricated conditions [31]. They showed that than with others FMs.
deformation is higher under dry condition than lubricating with Ringsberg made twin disc tests on R260 with 1% creepage and
water, and absent with grease, i.e., there is no anisotropy by 1.5 GPa in dry condition [47], for 5,000–10,000 cycles the crack
S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250 247

Fig. 9. EDS mapping of a longitudinal section of a rail sample; inset the EDS analysis of the MnS inclusion.

Fig. 10. BSE images of longitudinal section of the rail samples, 4,000 dry cycles followed by: a) 5,000, b) 10,000, and c) 20,000 wet cycles with HLK.
248 S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250

Fig. 11. BSE images of longitudinal section of the rail, a) 9,000, b) 14,000, and c) 24,000 dry cycles.

Fig. 12. a) Crack depth, and b) crack angle versus the fatigue cycles.

angle was 13–20°, which is high compared to 5° obtained at 9,000 wheel passes were 39° and 48° respectively [48]. In all cases,
cycles on R370CrHT. Although the tests conditions were different, cracks appear to grow faster when a fluid is added to preexisting
the fact that the cracks are shallower in R370CrHT than those cracks.
found by the author, means higher Rolling Contact Fatigue re-
sistance. Moreover, shallow cracks can cause lower mass losses
since the flakes detached from the surface will be smaller. How- 4. Conclusions
ever, more tests at the same conditions comparing both materials
are needed for further analysis. Eadie et al. carried out tests on a Dry and lubricated tests were carried out to evaluate the per-
full scale wheel–rail rig under dry conditions and using a com- formance of a friction modifier (FM) on preexisting cracks and its
s
mercial FM (KELTRACK HiRail), after 100,000 passes the average effects on the wear rate and the wear mechanisms of an R370CrHT
crack angle for the dry condition and applying FM every 500 fully pearlitic rail steel.
S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250 249

 The FM, Sintono Terra HLK, was able to maintain the friction at [12] D.F. Cannon, K.-O. Edel, S.L. Grassie, K. Sawley, Rail defects: an overview, Fa-
the expected levels of twin-disk tests, i.e., COT o 0.1. tigue Fract. Eng. Mater. Struct. 26 (2003) 865–886, http://dx.doi.org/10.1046/
 With 5% of creepage, the addition of a FM in presence of pre-
j.1460-2695.2003.00693.x.
[13] A. Ekberg, E. Kabo, Fatigue of railway wheels and rails under rolling contact
existing cracks increases the mass loss in the rail in comparison and thermal loading—an overview, Wear 258 (2005) 1288–1300, http://dx.
to the dry tests. However, combining the mass loss of the wheel doi.org/10.1016/j.wear.2004.03.039.
[14] U. Zerbst, K. Mädler, H. Hintze, Fracture mechanics in railway applications––an
and the rail up to 24,000 fatigue cycles showed that there is still overview, Eng. Fract. Mech. 72 (2005) 163–194, http://dx.doi.org/10.1016/j.
greater mass loss in the dry tests. engfracmech.2003.11.010.
 The rail wear rate (mass loss per cycle) in dry tests was almost [15] INNOTRACK, D4.4.1 – Rail Inspection Technologies〈http://www.innotrack.net/
IMG/pdf/d441.pdf〉, 2008.
constant for all fatigue cycles in these experiments, while for
[16] K. Sawley, J. Kristan, Development of bainitic rail steels with potential re-
the tests with FM the wear rate increased between 14,000 and sistance to rolling contact fatigue, Fatigue Fract. Eng. Mater. Struct. 26 (2003)
24,000 cycles due to flaking. 1019–1029, http://dx.doi.org/10.1046/j.1460-2695.2003.00671.x.
 For the evaluated conditions, no cracks on the surface of the [17] R. Lewis, U. Olofsson, Wheel-rail Interface Handbook, 1st ed, CRC Press;
Woodhead Pub, Boca Raton; Oxford, 2009.
wheel samples were observed. [18] S. Iwnicki, Handbook of Railway Vehicle Dynamics, 1 edition, CRC Press, Boca
 In the evaluated conditions, the wheel loses mass faster than Raton, 2006.
the rail and the magnitude of the adhesion and abrasion wear [19] European standard, EN 13674-1, - Railway applications - Track - Rail - Part 1:
Vignole railway rails 46 kg/m and above, Httpwwwen-Stand. (n.d.). 〈http://
mechanisms are significantly higher than the growth of cracks www.en-standard.eu/csn-en-13674-1-railway-applications-track-rail-part-1-
by fatigue. vignole-railway-rails-46-kg-m-and-above/〉 (accessed 30.08.13), 2011.
 The depth and angle of the cracks increase with the fatigue [20] European standard, EN 13262:2004 þ A1:2008 - Railway applications -
Wheelsets and bogies - Wheels - Product requirements, Httpwwwen-Stand.
cycles for both dry and lubricated conditions. For the dry sam- (n.d.). 〈http://www.en-standard.eu/〉 (accessed 04.11.13).
ples, between 14,000 and 24,000 fatigue cycles the crack angle [21] S. Maya-Johnson, Estudio del Comportamiento de Concentradores de Esfuerzo
seems to stabilize around 10°. Bajo Condiciones de Fatiga en Aceros para Rieles, (Master's Thesis), National
 The ductile inclusions follow the plastic deformation of the University of Colombia, 2014 〈http://www.bdigital.unal.edu.co/12071/〉.
[22] D.I. Fletcher, J.H. Beynon, Development of a machine for closely controlled
microstructure, become aligned with the crack tips, acting as rolling contact fatigue and wear testing, J. Test. Eval. 28 (2000) 267, http://dx.
stress concentrators and increasing the crack growth. doi.org/10.1520/JTE12104J.
[23] D.I. Fletcher, J.H. Beynon, Equilibrium of crack growth and wear rates during
unlubricated rolling-sliding contact of pearlitic rail steel, Proc. Inst. Mech. Eng.
F J. Rail Rapid Transit. 214 (2000) 93–105, http://dx.doi.org/10.1243/
Acknowledgements 0954409001531360.
[24] S.M.J. Felipe, Development of a lubrication system for wear and friction con-
trol in wheel/rail interfaces, (Doctoral Thesis), National University of Colombia,
The authors wish to acknowledge the support of the Medellín 2013.
City Hall with his program “Enlazamundos 2012”, co-supported by [25] W.R. Tyfour, J.H. Beynon, A. Kapoor, The steady state wear behaviour of
pearlitic rail steel under dry rolling-sliding contact conditions, Wear 180
Colciencias and the Metro system of Medellín-Colombia; contract
(1995) 79–89, http://dx.doi.org/10.1016/0043-1648(94)06533-0.
No 111850227421. Also thank to the LME at the Brazilian Nano- [26] C. Hardwick, R. Lewis, U. Olofsson, Low adhesion due to oxide formation in the
technology National Laboratory (LNNano). Research supported by presence of salt, Proc. Inst. Mech. Eng. F J. Rail Rapid Transit. 228 (2014)
LNNano - Brazilian Nanotechnology National Laboratory, CNPEM/ 887–897, http://dx.doi.org/10.1177/0954409713495666.
[27] Y.A. Areiza, S.I. Garcés, J.F. Santa, G. Vargas, A. Toro, Field measurement of
MCTI. coefficient of friction in rails using a hand-pushed tribometer, Tribol. Int. 82
(2015) 274–279, http://dx.doi.org/10.1016/j.triboint.2014.08.009.
[28] H. Harrison, T. McCanney, J. Cotter, Recent developments in coefficient of
friction measurements at the rail/wheel interface, Wear 253 (2002) 114–123,
References http://dx.doi.org/10.1016/S0043-1648(02)00090-X.
[29] J. Lundberg, M. Rantatalo, C. Wanhainen, J. Casselgren, Measurements of
friction coefficients between rails lubricated with a friction modifier and the
[1] A.F. Bower, The influence of crack face friction and trapped fluid on surface
wheels of an IORE locomotive during real working conditions, Wear 324–325
initiated rolling contact fatigue cracks, J. Tribol. 110 (1988) 704–711, http://dx. (2015) 109–117, http://dx.doi.org/10.1016/j.wear.2014.12.002.
doi.org/10.1115/1.3261717. [30] E.A. Gallardo-Hernandez, R. Lewis, Twin disc assessment of wheel/rail adhe-
[2] M. Kaneta, H. Yatsuzuka, Y. Murakami, Mechanism of crack growth in lu- sion, Wear 265 (2008) 1309–1316, http://dx.doi.org/10.1016/j.
bricated rolling/sliding contact, Tribol. Trans. 28 (1985) 407–414, http://dx.doi. wear.2008.03.020.
org/10.1080/05698198508981637. [31] C. Hardwick, R. Lewis, D.T. Eadie, Wheel and rail wear—Understanding the
[3] Y. Murakami, M. Kaneta, H. Yatsuzuka, Analysis of surface crack propagation in effects of water and grease, Wear 314 (2014) 198–204, http://dx.doi.org/
lubricated rolling contact, Tribol. Trans. 28 (1985) 60–68, http://dx.doi.org/ 10.1016/j.wear.2013.11.020.
10.1080/05698198508981595. [32] S.R. Lewis, R. Lewis, G. Evans, L.E. Buckley-Johnstone, Assessment of railway
[4] M. Kaneta, Y. Murakami, Effects of oil hydraulic pressure on surface crack curve lubricant performance using a twin-disc tester, Wear 314 (2014)
growth in rolling/sliding contact, Tribol. Int. 20 (1987) 210–217, http://dx.doi. 205–212, http://dx.doi.org/10.1016/j.wear.2013.11.033.
org/10.1016/0301-679X(87)90076-4. [33] U. Olofsson, R. Nilsson, Surface cracks and wear of rail: a full-scale test on a
[5] V. Mota, L.A. Ferreira, Influence of grease composition on rolling contact wear: commuter train track, Proc. Inst. Mech. Eng. F J. Rail Rapid Transit. 216 (2002)
experimental study, Tribol. Int. 42 (2009) 569–574, http://dx.doi.org/10.1016/j. 249–264, http://dx.doi.org/10.1243/095440902321029208.
triboint.2008.04.002. [34] U. Zerbst, R. Lundén, K.-O. Edel, R.A. Smith, Introduction to the damage tol-
[6] J. Sundh, U. Olofsson, K. Sundvall, Seizure and wear rate testing of wheel–rail erance behaviour of railway rails – a review, Eng. Fract. Mech. 76 (2009)
contacts under lubricated conditions using pin-on-disc methodology, Wear 2563–2601, http://dx.doi.org/10.1016/j.engfracmech.2009.09.003.
265 (2008) 1425–1430, http://dx.doi.org/10.1016/j.wear.2008.03.025. [35] R. Stock, D.T. Eadie, D. Elvidge, K. Oldknow, Influencing rolling contact fatigue
[7] S. Bogdański, P. Lewicki, M. Szymaniak, Experimental and theoretical in- through top of rail friction modifier application – A full scale wheel–rail test
vestigation of the phenomenon of filling the RCF crack with liquid, Wear 258 rig study, Wear 271 (2011) 134–142, http://dx.doi.org/10.1016/j.
(2005) 1280–1287, http://dx.doi.org/10.1016/j.wear.2004.03.038. wear.2010.10.006.
[8] W.R. Tyfour, J.H. Beynon, A. Kapoor, Deterioration of rolling contact fatigue life [36] D.I. Fletcher, P. Hyde, A. Kapoor, Modelling and full-scale trials to investigate
of pearlitic rail steel due to dry-wet rolling-sliding line contact, Wear 197 fluid pressurisation of rolling contact fatigue cracks, Wear 265 (2008)
(1996) 255–265, http://dx.doi.org/10.1016/0043-1648(96)06978-5. 1317–1324, http://dx.doi.org/10.1016/j.wear.2008.02.025.
[9] D.I. Fletcher, J.H. Beynon, The influence of lubricant type on rolling contact [37] J.H. Beynon, J.E. Garnham, K.J. Sawley, Rolling contact fatigue of three pearlitic
fatigue of pearlitic rail steel, in: M.P.D. Dowson (Ed.), Tribol. Ser., Elsevier, rail steels, Wear 192 (1996) 94–111, http://dx.doi.org/10.1016/0043-1648(95)
1999, 06776-0.
pp. 299–310 〈http://www.sciencedirect.com/science/article/pii/ [38] J.E. Garnham, J.H. Beynon, Dry rolling-sliding wear of bainitic and pearlitic
S0167892299800510〉 (accessed 08.10.13). steels, Wear 157 (1992) 81–109, http://dx.doi.org/10.1016/0043-1648(92)
[10] V. Reddy, G. Chattopadhyay, P.-O. Larsson-Kråik, D.J. Hargreaves, Modelling 90189-F.
and analysis of rail maintenance cost, Int. J. Prod. Econ. 105 (2007) 475–482, [39] S. Bogdañski, A rolling contact fatigue crack driven by squeeze fluid film, Fa-
http://dx.doi.org/10.1016/j.ijpe.2006.03.008. tigue Fract. Eng. Mater. Struct. 25 (2002) 1061–1071, http://dx.doi.org/10.1046/
[11] L. Roger, Survey of wheel/rail lubrication practices〈http://www.rssb.co.uk/RE j.1460-2695.2000.00563.x.
SEARCH/Lists/DispForm_Custom.aspx?ID ¼ 29〉 (accessed 08.10.13), 2003. [40] J.E. Garnham, C.L. Davis, The role of deformed rail microstructure on rolling
250 S. Maya-Johnson et al. / Wear 380-381 (2017) 240–250

contact fatigue initiation, Wear 265 (2008) 1363–1372, http://dx.doi.org/ propagation in pearlitic rail steel, Wear 314 (2014) 57–68, http://dx.doi.org/
10.1016/j.wear.2008.02.042. 10.1016/j.wear.2013.11.034.
[41] J.E. Garnham, C.L. Davis, Very early stage rolling contact fatigue crack growth [46] D.I. Fletcher, J.H. Beynon, The effect of intermittent lubrication on the fatigue
in pearlitic rail steels, Wear 271 (2011) 100–112, http://dx.doi.org/10.1016/j. life of pearlitic rail steel in rolling-sliding contact, Proc. Inst. Mech. Eng. F J.
wear.2010.10.004. Rail Rapid Transit. 214 (2000) 145–158, http://dx.doi.org/10.1243/
[42] A.D. Wilson, Fatigue crack propagation in A533B steels, J. Press. Vessel Tech- 0954409001531270.
nol. 99 (1977) 459–469, http://dx.doi.org/10.1115/1.3454560. [47] J.W. Ringsberg, Life prediction of rolling contact fatigue crack initiation, Int.
[43] S. Ganesh Sundara Raman, Y. Kitsunai, Fatigue crack growth behaviour of J. Fatigue. 23 (2001) 575–586, http://dx.doi.org/10.1016/S0142-1123(01)
A533B steel in pressurized water at 288 and 28 °C, Eng. Fail. Anal. 11 (2004) 00024-X.
293–303, http://dx.doi.org/10.1016/j.engfailanal.2003.08.009. [48] D.T. Eadie, D. Elvidge, K. Oldknow, R. Stock, P. Pointner, J. Kalousek, P. Klauser,
[44] S. Maya-Johnson, A.J. Ramirez, A. Toro, Fatigue crack growth rate of two
The effects of top of rail friction modifier on wear and rolling contact fatigue:
pearlitic rail steels, Eng. Fract. Mech. 138 (2015) 63–72, http://dx.doi.org/
full-scale rail–wheel test rig evaluation, analysis and modelling, Wear 265
10.1016/j.engfracmech.2015.03.023.
(2008) 1222–1230, http://dx.doi.org/10.1016/j.wear.2008.02.029.
[45] N. Larijani, J. Brouzoulis, M. Schilke, M. Ekh, The effect of anisotropy on crack

View publication stats

Anda mungkin juga menyukai