Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Science & Engineering A 722 (2018) 147–155

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Role of crystallographic orientation and grain boundaries in fatigue crack T


propagation in used pearlitic rail steel

Mohammad Masoumia, , Edwan Anderson Arizaa,b, Amilton Sinatorac,d, Helio Goldensteina
a
Department of Metallurgical and Materials Engineering, University of São Paulo, Av. Prof. Mello Moraes, 2463, São Paulo, SP CEP 05508-030, Brazil
b
Mechanical Technology Program, Technological University of Pereira, Vereda la Julita, Pereira, Risaralda, Colombia
c
Department of Mechanical Engineering, University of São Paulo, Av. Prof. Mello Moraes 2231, 05508-970 São Paulo, SP CEP 05508-900, Brazil
d
Instituto Tecnológico Vale, Av. Juscelino Kubitschek 31, Bauxita, 35400-000 Ouro Preto, MG, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: The possible effects of crystallographic orientation and grain boundary distributions on fatigue crack initiation
Crystallographic orientation and propagation were studied in used pearlitic rail steels. Microstructure and hardness variation along the depth
Taylor factor distance showed the effect of strain accumulation on microstructure modification. Although no evidence was
Kernel average misorientation shown of a correlation of inclusions and oxides with crack initiation, a high-pressure water environment entering
into a crack can accelerate its propagation. Local crystallographic orientations analyzed closed to the crack path
revealed that {110} grains oriented along the train passage plane, with a high Taylor factor and kernel average
misorientation, are highly prone to crack propagation. This could be explained by the effect of local stress
concentrations due to persistent slip bands.

1. Introduction the high thermal conductivity of the rail, martensite forms by rapid
cooling. Recently, researchers reported that the cementite dissolution of
Pearlitic steels are representative of the strongest steels, and have pearlite by severe deformation and carbon decoration at ferrite grain
been widely used in railroad rails and engineering springs because of boundaries is responsible for WEL formation as a similar mechanism to
their high wear resistance and strength [1,2]. It is well known that the that responsible for cementite dissolution of cold-drawn pearlitic steel
strength of pearlitic steels is associated with a Hall–Petch type re- wires [11–14]. The hard, brittle, and featureless WEL structure formed
lationship with respect to the interlamellar spacing [3,4]. This is ex- at regions close to the running contact surface can increase the wear
plained by a dislocation pile-up model to restrict the dislocation mo- rate. A large amount of surface shear deformation due to the finite
bility. On the other hand, the residual stress, generated from the misfit contact patch size, together with high levels of stress and friction lead to
plastic strain between cementite and ferrite, increased with the lamellar the formation of microcracks, significantly increasing the wear and
spacing, can reduce the toughness. In a heavy-haul railway, the rail and RCF, which may cause early fracture. The development of heavy-haul
wheels are subjected to intense plastic deformation from vertical, tan- railways imposes a considerable materials challenge in terms of the
gential, and longitudinal loads [5,6]. Under conditions where high microstructural stability and wear resistance of the surfaces of rails
contact loads are in action, elastic shakedown is reached, giving ben- [5,15,16]. The aim of this work was to explore the potential effects of
eficial work hardening and compressive stresses, and enhancing the the crystallographic orientation and grain boundary distributions on
monotonic and cyclic performance in the material. An excess of surface the initiation and propagation of fatigue cracks. Initially, micro-
shear stresses may cause rolling contact fatigue (RCF) damage from the structural changes along the depth distance from the surface were
accumulation of plastic strains [6–8], which results in the formation of studied. Then, detailed studies using the electron backscattered dif-
a modified surface structure called a white etching layer (WEL). It has fraction technique (EBSD) were performed to analyze the relation be-
been assumed that heat is generated at the contact zone due to friction tween the crystallographic orientation and prevention or enhancement
during wheel–rail rolling/sliding contact. It is reported that the tem- of crack propagation.
perature at the contact region can reach 700 °C above the eutectic point
due to the energy dissipated in the contact patch associated with lo-
calization of adiabatic shear deformation [9,10]. Consequently, due to


Corresponding author.
E-mail address: mohammad.masoumi@usp.br (M. Masoumi).

https://doi.org/10.1016/j.msea.2018.03.028
Received 15 January 2018; Received in revised form 5 March 2018; Accepted 6 March 2018
Available online 07 March 2018
0921-5093/ © 2018 Elsevier B.V. All rights reserved.
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

Fig. 1. (a) Photograph of railway rail showing the head check defects concentrated at gauge corner of the head crown, (b) SEM micrographs indicating the crack fatigue path across cross-
section of the railhead.

2. Material and experimental procedures distance from the running surface. Three categories of pearlite
morphologies, namely featureless, deformed (i.e., fibrous and curly
A used pearlitic rail of near-eutectoid steel with head check defects lamellar), and non-deformed pearlite structure, were characterized as a
(Fig. 1) was removed from a 344 m radius curve after approximately 30 function of depth distance from the railhead surface. Fig. 2b shows the
MTB freight had passed over it. The analyzed chemical composition of heavily deformed superficial featureless WEL at about 30 µm depth
the rail is listed in Table 1. from the surface. Fig. 3 also shows the presence of a WEL region close to
The microstructure of the used rail sample was investigated using a the rail surface, associated with a high etching resistance to the Nital
scanning electron microscope (SEM, FEI-Inspect F50) operating at etchant. The non-uniform carbon content in the solute solution and/or
20 kV. Before the investigations a sample was cut off from the cross- differences in the content and size of the carbide particles can cause the
section of the crown head. The sample surface was ground with SiC formation of non-conventional supersaturated ferrite (WEL structure).
paper from 100 to 2500 grit, followed by polishing using 6, 3, and 1 µm A combination of the normal load (weight) and tractions stresses gen-
diamond suspensions. Finally, the sample was etched with a 2% solu- erated by the vehicle steering forces in the contact surface gives rise to
tion of nitric acid in ethanol (Nital) for 20 s. Electron backscattered shear stresses in the rails [18]. The intense shear and the compressive
diffraction (EBSD) analysis was performed to investigate the local stresses introduced by train passage reduce the interlamellar spacing
crystal orientation and boundary distributions. Before the EBSD test, (i.e. ferrite – cementite) and reorients the pearlite morphology parallel
the samples were prepared according to the standard preparation and to the axis of train passage. The extremely high dislocation densities at
polished with 50 nm colloidal silica slurry for 3 h. Kikuchi patterns were the WEL generated in the contact rail surface provoke cementite dis-
acquired by using a TSL EDAX system installed on a FEI-SEM with solution, leading to a featureless structure. The required energy for
20 kV, a working distance of about 12 mm and 50 nm step size. Then, cementite dissolution is provided by two possible sources: i) the for-
the data was processed and displayed using TSL OIM analysis 7. It is mation of an additional area of cementite/ferrite boundaries as a result
notable that EBSD was not able to detect very fine cementite particles of shear deformation [19,20] and ii) trapping of the carbon atoms by
within the pearlitic lamellae [17]. In addition, the variations in hard- dislocations that pile up at the ferrite/cementite phase boundaries
ness through the depth of the rail were measured by a Shimadzu (HMV- [21,22]. In other words, the intense plastic deformation at the rail
2TADW) microhardness tester with a load of 2.97 N (300 g) for 15 s surface introduces a large number of dislocations so that the carbon
(HV0.3). atoms dissolve from the cementite and then preferentially sit at the
dislocation walls in the ferrite structure, relieving the strain on the
lattice. Fig. 2c shows the distorted pearlite structure with a fibrous la-
3. Results and discussion
mellar area below the superficial WEL about 50 µm below the railhead
surface. It was shown that the pearlite lamellae became thinner and
Fig. 2 shows the changes of pearlite morphology as a function of
aligned along the train passage direction by increasing the loading load
and service time. Embury and Fisher [23] reported that the fibre
Table 1
Chemical composition obtained by emission spectrometry technique of the tested rail
structure acts as a strong barrier to dislocation motion, leading to sig-
steel (in wt%). nificant increases in strengthening. The non-deformed, as-hot rolled
pearlite morphology located 1 cm below the contact surface is pre-
C Si Mn P S Al Cr Ni sented in Fig. 2d. Fine pearlite with no preferred colonies orientation
0.75 0.22 1.03 0.019 0.005 0.002 0.21 0.05
was observed in this region, which is attributed to the fabrication and
heat treatment processes.

148
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

Fig. 2. (a) Photograph of railway rail sample, (b) superficial WEL, (c) distorted and realigned pearlite morphology at ≈ 50 µm from the surface, and (d) non-deformed pearlite
morphology located at 1 cm below the contact surface.

Fig. 3. Microstructural features visible using optical microscopy, etched by 2% nital so-
lution.

Fig. 4 shows the variation of Vickers microhardness along the depth Fig. 4. Variation of Vickers microhardness as a function of distance from the contact
distance. The results showed that the hardness decreased with in- surface.
creasing depth distance. The superficial WEL subjected to intense
plastic deformation had the highest hardness value (approximately HV0.3. This was attributed to the maximum shear and compression
574 ± 25 HV0.3). Then, a continuous reduction of hardness was ob- stresses concentrated on the surface, precipitation hardening from
served at ≈ 6 mm from the surface, reaching the value of 375 ± 5 broken cementite particles, and work hardening at the rail surface.

149
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

water drawn into a surface microcrack, pressurized by a passing wheel,


can accelerate crack growth to a greater depth.
The EBSD technique is widely used to identify the crystal orienta-
tion and measure the local strain by means of diffraction criteria re-
sulting from Kikuchi bands, providing an extensive set of crystal-
lographic data to explain the reason for crack formation. Orientation
image maps (OIM) in different areas at 0.5 mm depth from the running
rail surface are shown in Fig. 7. Colour coded OIM present the variation
of crystallographic orientation parallel to the train's running direction
(TD).
The mutual interaction of the rail and wheel in service generates
huge dislocation densities in the contact surface. The dislocation walls
and sub-grains absorb the dislocations to reduce the internal energy,
leading to the formation of a new set of ultrafine grains. Moreover, the
rotation or reorientation of subgrains during deformation generates a
high fraction of HABs (high angle boundaries). Fig. 8 shows the var-
iation of average grain sizes and misorientation angles of neighbouring
points obtained from the related ESBD data. It is notable that average
grain sizes were determined by the threshold misorientation angle of
Fig. 5. SEM micrograph of bent and fragments pearlite morphology close the micro-
defects closed the surface. 15° of neighbouring crystals. The average ferrite matrix grain size of the
as-received sample was 11 ± 1 µm. Thus, a significant grain refine-
ment occurred under intense plastic deformation due to the above-
Then, a gradual hardness reduction was observed as a function of the
mentioned mechanisms, as shown in Fig. 8. However, the cracked re-
contact patch length [24]. The hardness results are in agreement with
gions showed slightly larger grain sizes (≈ 1.6 ± 0.3 µm), accom-
the mentioned tendency. It is worth noting that the higher hardness and
panied with higher average misorientation angles (≈ 25° ± 2). The
compressive residual stress at the WEL can increase the wear resistance.
stored energy of plastic deformation can be estimated from the dis-
However, superficial microcracks can grow and easily propagate in the
location boundary spacing and misorientation angles. According to the
highly deformed rail surface because of its brittle nature, leading to
Read–Shockley equation [30], the energy per unit area of the disloca-
early fracture. For this reason, rail companies spend billions of dollars
tion boundary is directly related to the misorientation angles. Thus, a
per year to remove superficial WEL and microcracks by grinding.
higher amount of stored energy due to high average misorientation
The formation of microcracks and debris at the rail surface is in-
angles can indicate the localized strain atomic misfit energy to accel-
evitable due to the interaction of the wheels and rail. Fig. 5 shows the
erate the crystal failure.
microstructural changes (i.e., bending and fragments of pearlite la-
Material properties vary with the crystal orientation in poly-
mellae) close to the subsurface microdefects (approximately 50 µm
crystalline materials. The sequence of the orientation dependence of
from the surface). According to the Hertzian contact stress theory [24],
yield stress (YS) in a BCC structure is YS111 > YS112 > YS110 > YS100 ,
the shear stress reaches the maximum values below the contact surface
which is associated with the crystallographic interplanar spacing and
and could exceed the elastic limit. Head checks and shelling problems
surface energy of each specific plane [31–33]. Fig. 9 shows the inverse
reported by Brazilian railroads caused by the surface cracks are origi-
pole figure (IPF) parallel to the train passage direction, calculated from
nated by the subsurface shearing stresses [25]. Toribio et al. [26] and
related OIMs to characterize the distribution of local crystallographic
Guo et al. [27] explained the complicated wavy and curly pearlite
orientations based on orthorhombic crystal system symmetry with re-
morphologies by rotational movements, when the two ends of pearlite
spect to the crystallographic axes. Fig. 9 also shows the normalized
lamellae received the same compressive stress perpendicular to the
numerical fraction of some common crystallographic textures with a
rolling direction. Cementite lamellae parallel to or with a small angle to
deviation less than 2.5°. It was found that grains oriented along {111}
the train passage direction are lengthened along the rolling direction
crystallographic axes parallel to the train passage direction were pre-
and thinned in the process. Kink or shear bands also enhance the
dominant in non-cracked OIMs, while {110} grains parallel to the train
fragmenting and reorientation of the cementite layers. In addition, on
passage direction were dominant in cracked regions. Non-compact
pearlite lamellae with a moderate angle to the train passage direction,
{100} planes with the largest interatomic spacing have preferred sites
rotation is predominant rather than thinning. Therefore, curved shape
to initiate crystallographic micro-defects. Compact slip {110} planes
and wavy cementite particles are developed by increasing the strain
accelerated the propagation by dislocation motion/multiplication. Al-
induced due to the friction between the wheels and rails.
though no evidence of micro-crack initiation was found along {001}
Head checks and rolling contact fatigue cracks, which are known as
planes, the {110} planes were predominant in the vicinity of the crack
a major source of failure in the heavy-haul Brazilian railroad, were
path. The initiation and propagation of fatigue cracks were divided into
observed in the investigated sample. Because of the improvement of
four main categories by Mikulich et al. [34], i.e., crack growth along: i)
quality control in modern rail production, very fine and negligible in-
grain boundaries, ii) twin boundaries, iii) slip bands, and iv) transgra-
clusions (essentially, alumina- and silica-based) were observed. For
nular crack propagation along grains with low-indexed planes such as
example, through continuous casting processing, the newly smelted
{001} planes. Therefore, {110} planes are known as the primary slip
iron from the blast furnace is taken in the liquid from the convertor by
planes in BCC structure, leading to the activation of multiple slip sys-
employing ladles directly into blooms. This process eliminates casting
tems for a given grain, resulting in an accelerated rate of crack growth.
defects such as cracks, hot tears, inclusion, and blow holes [28]. No
The Taylor factor can predict the level of resistance to deformation
evidence was found to show any correlation between the oxides and
at a particular point and the crystal orientation for a specific stress state
crack initiation or propagation. In addition, energy dispersive X-ray
surrounded by neighbouring crystals [35]. Deformation in a poly-
spectroscopy (EDX) analyses characterized the dispersion of most
crystalline material depends on many factors such as the distribution of
oxides in the crack path (such as Mn, Al, and Si), as observed in Fig. 6.
crystal orientation, strain hardening, stress state, and anisotropy. Ac-
Environmental conditions such as temperature, humidity, train speed,
cording to the von Mises criterion, at least five sets of slip systems are
track geometry, and wheel and rail profiles can cause the formation of
required for plastic deformation to occur. Slip occurs on specific crys-
oxides particles in the crack path. Bower et al. [29] explained that the
tallographic planes and specific crystallographic directions with high

150
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

Fig. 6. Dispersion of oxides in crack path characterized by EDX.

Fig. 7. OIM maps for different regions at the same distance from the surface: (a, b) non-cracked and (c, d) closed to the cracked regions.

151
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

systems (i.e. planes and directions) require the minimum expenditure


energy for plastic deformation (Taylor factor < 1, blue and light green
grains). Second are moderate Taylor factor grains, which can slip with
appropriate rotation (1 < Taylor factor < 3, dark green and yellow
grains). And finally, high Taylor factor or hard grains, which are not
able to rotate and to meet appropriate slip systems, are highly prone to
crack formation (3 < Taylor factor, red grains). The results showed
that the low and intermediate Taylor factors belonged the OIM a and b
(referring to Fig. 7), while the high level Taylor factor (about 55%) at
the cracked regions (Fig. 10c and d) indicates the regions highly prone
to failure. Therefore, the lack of enough slip systems restricted the
dislocation movements, leading to increasing the local stress con-
centration by dislocation pile-ups. The local stress concentration
around the crack and crystal defects promote the formation of an
asymmetric lattice disclination and enhance the crystallographic de-
fects and lattice distortion.
Fig. 8. Variation of average grain sizes and misorientation angles of different OIMs (re-
Fig. 11a shows the distribution of boundary types of ferrite matrix in
ferring to Fig. 7).
the pearlite structure in three regions, non-deformed (as-received) and
close to the contact surface, including: i) non-cracked and ii) cracked
regions. This indicates that the severe plastic deformation close to the
running surface changed the non-deformed structure with a high frac-
tion of LABs (low angle boundaries) into a wavy and fibrous structure
with a large number of HABs. In other words, the high storage energy
due to dislocation interaction and lattice distortion are responsible for
the formation of a new set of ultrafine grains with a high fraction of
HABs. In addition, to find an explanation for the crack formation close
to the running surface, the kernel average misorientation (KAM) was
calculated from the first five nearest neighbours with a 5° maximum
misorientation limit [36] to represent the dislocation densities in in-
dividual grains, as shown in Fig. 11b. The low KAM values (less than
2.5°) of the non-deformed and non-defective zones close to the running
surface indicated the annihilation of the dislocations and the formation
of new low-strain grains, while high KAM values (greater than 2.5°),
because of highly distorted grains with high storage energy, led to a
reduction in local ductility, facilitating crack formation.
Fig. 9. Inverse pole figures of each OIMs analysis and normalized numerical fraction of The mutual interaction between the microstructural features (i.e.,
main crystallographic texture (referring to Fig. 7). microstructure and grain orientation), extrinsic extension ahead of the
crack tip, and residual plastic deformation caused by crack closure ef-
atomic packing density. In BCC structure, {011} < 11U fects all played a significant role in the crack growth behaviour under
+012B > (considered as the primary slip system) was used in Taylor cyclic loading by train passage [37–39]. The crack closure area is
factor analyses (Fig. 10). According to the Taylor theory, the grains can considered as a compressive residual stress zone that is left around the
be categorized into three groups, depending on the minimum energy of crack path. It is well known that the formation of microcracks and/or
slip. First, low Taylor factor grains, which are already aligned into slip voids in the plastic region ahead of the crack tip leads to propagation of

Fig. 10. (a-d) Taylor factor maps, and (e) distribution of Taylor factor in different regions. The “a” to “d” identification is related to Fig. 7.

152
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

Fig. 11. Distribution of: (a) boundary types and (b) KAM values in three different regions: non-deformed, and closed the contact surface including non-cracked, and cracked regions.

Fig. 12. Schematic illustration of mutual interaction between intrinsic and extrinsic mechanisms of crack-tip, measured EBSD regarding each region, and variation of misorientation angle
at perpendicular direction of crack propagation.

the crack by cleavage, intergranular cracking, and microvoid coales- affected plastic zone. As a summary, the strain energy induced by dy-
cence mechanisms. The effects of lattice distortion due to residual namic rotation recrystallization generates a new set of ultrafine grains
plastic deformation on the variation of dislocation densities and lattice in the vicinity of the crack path. In addition, the crystallographic or-
distortion in the perpendicular direction of crack propagation are re- ientations can promote the generation of dislocations, providing ade-
presented in Fig. 12. The local stress concentration around the crack quate slip systems and dislocation multiplications. The effect of local
and crystal defects promote the formation of an asymmetric lattice stress concentrations by persistent slip bands accelerates the rate of
disclination and enhance lattice distortion. Consequently, the formation crack growth, as reported by Brown [40], which is in agreement with
of ultra-fine grains with a higher number of HABs leads to the increase the results presented here, where the main slip {110} systems were
of the average misorientation to about 55° in this zone. Misorientation predominant around the cracked regions.
between neighbouring points indicates the mismatch and lattice dis- Under cyclic deformation due to train passage, lattice mismatches in
tortion, made up of arrays of dislocations due to lattice rotation during the different crystalline structure from one grain to another hinder the
the cyclic or tensile deformation. This is because the misorientation dislocation movements; hence, the dislocation pile-ups are created at
angles, increased by dislocation pile-up to higher than 15°, result in the random high grain boundaries. The mentioned phenomena, besides the
formation of ultrafine grains. In general, the grain boundary energy and dislocation multiplication and dislocation–dislocation interactions on
dislocation densities increased around the crack path due to the the slip bands, enhanced intergranular crack propagation in the

153
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

material [41,42]. Since a further cycling deformation intensified the and Rail Interface Issues, International Heavy Haul Association, Virginia, 2001.
lattice distortion from the boundaries into the interior of the grains, [6] K. Cvetkovski, J. Ahlström, Characterisation of plastic deformation and thermal
softening of the surface layer of railway passenger wheel treads, Wear 300 (2013)
transgranular crack propagation can occur. The tendency of boundary 200–204, http://dx.doi.org/10.1016/j.wear.2013.01.094.
obstruction effects at random boundaries leads to intergranular fatigue [7] A. Kapoor, D.I. Fletcher, F.J. Franklin, The role of wear in enhancing rail life, tri-
crack nucleation due to the high stress concentrations at the HABs. The bological research and design for engineering systems, in: Proceedings of the 29th
Leeds-Lyon Symposium on Tribology, 41, 2003, pp. 331–340. 〈http://dx.doi.org/
intergranular crack propagation occurs in the high Taylor factor grains, dx.doi.org/10.1016/S0167-8922(03)80146-3〉.
while the transgranular crack propagation is characterized at the grain [8] R.L. Reuben, Monitoring of rail-wheel interaction using acoustic emission (AE),
boundaries with a high Taylor factor mismatch due to the formation of Adv. Mater. Res. 13 (2006) 1–24, http://dx.doi.org/10.4028/www.scientific.net/
AMR.13-14.161.
shear bands and lattice distortions [43,44]. In addition, the combina- [9] T.S. Eyre, A. Baxter, The formation of white layers at rubbing surfaces, Tribology 5
tion of a large fraction of random boundaries associated with the (1972) 256–261, http://dx.doi.org/10.1016/0041-2678(72)90104-2.
dominant {110} planes parallel to the train passage direction, accom- [10] W. Lojkowski, M. Djahanbakhsh, G. Bürkle, S. Gierlotka, W. Zielinski, H.J. Fecht,
Nanostructure formation on the surface of railway tracks, Mater. Sci. Eng. A 303
panied by a high level of Taylor factor and KAM values (highly de-
(2001) 197–208, http://dx.doi.org/10.1016/S0921-5093(00)01947-X.
formed grains), could enhance the nucleation and propagation of fa- [11] F. Wetscher, R. Stock, R. Pippan, Changes in the mechanical properties of a pearlitic
tigue cracks. steel due to large shear deformation, Mater. Sci. Eng. A 445–446 (2007) 237–243,
http://dx.doi.org/10.1016/j.msea.2006.09.026.
[12] Y.J. Li, P. Choi, S. Goto, C. Borchers, D. Raabe, R. Kirchheim, Evolution of strength
4. Conclusions and microstructure during annealing of heavily cold-drawn 6.3 GPa hypereutectoid
pearlitic steel wire, Acta Mater. 60 (2012) 4005–4016, http://dx.doi.org/10.1016/
The potential effects of crystallographic orientation and grain j.actamat.2012.03.006.
[13] F. Fang, Y. Zhao, L. Zhou, X. Hu, Z. Xie, J. Jiang, Texture inheritance of cold drawn
boundary distributions on the initiation and propagation of rolling pearlite steel wires after austenitization, Mater. Sci. Eng. A 618 (2014) 505–510,
contact fatigue cracks in a used pearlitic rail steel were analyzed. The http://dx.doi.org/10.1016/j.msea.2014.09.061.
main conclusions are summarized as follows: [14] X. Zhang, A. Godfrey, X. Huang, N. Hansen, Q. Liu, Microstructure and strength-
ening mechanisms in cold-drawn pearlitic steel wire, Acta Mater. 59 (2011)
3422–3430, http://dx.doi.org/10.1016/j.actamat.2011.02.017.
• No evidence was found to show any correlation of inclusions and [15] P. Micenko, H. Li, Double dip hardness profiles in rail weld heat-affected zone —
literature and research review report, Lit. Res. Rev. Rep.: Improv. Railw. Weld.
oxides with crack initiation.
• Grains oriented along {110} planes parallel to the train passage (2013) 1–28.
[16] E.E. Magel, Rolling contact fatigue: a comprehensive review, Washington,
direction enhance crack propagation, while {111} grains showed DC20590, 2011.
more resistance to fatigue crack formation. [17] A. Durgaprasad, S. Giri, S. Lenka, S. Kundu, S. Mishra, S. Chandra, R.D. Doherty,

• The pearlite lamellae became thinner and aligned along the train I. Samajdar, Defining a relationship between pearlite morphology and ferrite
crystallographic orientation, Acta Mater. 129 (2017) 278–289.
passage direction with increasing loading and service time. Fine [18] A.D. Hearle, L. Johnson, T. Street, Mode II stress intensity factors for a crack parallel
pearlite with no preferred colonies orientation was observed to 1 cm to the surface of an elastic half- space subjected to a moving point load, J. Mech.
below the contact surface and was attributed to the non-deformed Phys. Solids 33 (1985) 61–81.
[19] S.B. Newcomb, W.M. Stobbs, A transmission electron microscope study of the white
as-fabricated process. etching layer on a rail head, Mater. Sci. Eng. A 66 (1984) 195–204.
• The intense shear and the compressive stresses introduced by train [20] M.H. Hong, W.T. Reynolds, T. Tarui, K. Hono, Atom probe and transmission elec-
tron microscopy investigations of heavily drawn pearlitic steel wire, Metall. Mater.
passage reduce the interlamellar spacing and reorient the pearlite
Trans. A 30 (1999) 717–727, http://dx.doi.org/10.1007/s11661-999-1003-y.
morphology. [21] Y. Wang, T. Lei, J. Liu, Tribo-metallographic behavior of high carbon steels in dry
• The extremely high dislocation densities generated in the superficial sliding, Wear 231 (1999) 12–19, http://dx.doi.org/10.1016/S0043-1648(99)
00116-7.
WEL provoked cementite dissolution, leading to a featureless WEL
[22] Y.J. Li, P. Choi, C. Borchers, S. Westerkamp, S. Goto, D. Raabe, R. Kirchheim,
structure. Atomic-scale mechanisms of deformation- induced cementite decomposition in
• The greater amount of high average misorientation angles and lat- pearlite, Acta Mater. 59 (2011) 3965–3977, http://dx.doi.org/10.1016/j.actamat.
2011.03.022.
tice distortion in the direction perpendicular to the crack path in-
[23] X. Zhang, A. Godfrey, N. Hansen, X. Huang, W. Liu, Q. Liu, Evolution of cementite
dicated the localized strain atomic misfit energy.

morphology in pearlitic steel wire during wet wire drawing, Mater. Charact. 61
Grains with a high Taylor factor and kernel average misorientation (2009) 65–72, http://dx.doi.org/10.1016/j.matchar.2009.10.007.
are highly prone to crack nucleation and propagation due to lattice [24] V.H.H. Hertz, Über die Berührung fester elastischer Körper, J. Die Reine Angew.
Math. 92 (1881) 156–171.
incompatibilities and highly distorted grains with high stored en-
[25] F.C. Santos, A.A. Santos Jr, F. Bruni, L.T. Santos, Evaluation of subsurface contact
ergy. stresses in railroad wheels using an elastic half-space model, J. Braz. Soc. Mech. Sci.
Eng. 26 (2004) 420–429, http://dx.doi.org/10.1590/S1678-58782004000400007.
Acknowledgments [26] J. Toribio, Relationship between microstructure and strength in eutectoid steels,
Mater. Sci. Eng. A 389 (2004) 227–230, http://dx.doi.org/10.1016/j.msea.2004.
01.084.
The authors are grateful to FAPESP, the Research Foundation of the [27] G. Ning, L. Baifeng, W. Bingshu, L. Qing, Microstructure and texture evolution in
State of São Paulo, and Vale Institute of Technology (ITV) for their fi- fully pearlitic steel during wire drawing, Sci. China Technol. Sci. 56 (2013)
1139–1146, http://dx.doi.org/10.1007/s11431-013-5184-7.
nancial support. [28] J.S. Mundrey, Railway Track Engineering, 4 edition, Tata McGraw-Hill Education,
2009.
References [29] A.F. Bower, The influence of crack face friction and trapped fluid on surface in-
itiated rolling contact fatigue cracks, Trans. Asme. 110 (1988) 704–711.
[30] W.T. Read, W. Shockley, Dislocation models of crystal grain boundaries, Phys. Rev.
[1] O.P. Modi, N. Deshmukh, D.P. Mondal, A.K. Jha, A.H. Yegneswaran, H.K. Khaira, 78 (1950) 1536–1552.
Effect of interlamellar spacing on the mechanical properties of 0.65% C steel, Mater. [31] D. De Knijf, T. Nguyen-minh, R.H. Petrov, L.A.I. Kestens, J. John, Orientation de-
Charact. 46 (2001) 347–352, http://dx.doi.org/10.1016/S1044-5803(00)00113-3. pendence of the martensite transformation in a quenched and partitioned steel
[2] M. Wu, L. Hua, Y.C. Shao, Q.J. Zhou, Influence of the annealing cooling rate on the subjected to uniaxial tension, J. Appl. Crystallogr. 47 (2014) 1261–1266, http://dx.
microstructure evolution and deformation behaviours in the cold ring rolling of doi.org/10.1107/S1600576714011959.
medium steel, Mater. Des. 32 (2011) 2292–2300, http://dx.doi.org/10.1016/j. [32] R. Blondé, E. Jimenez-Melero, L. Zhao, J.P. Wright, E. Brück, S. Van Der Zwaag,
matdes.2010.11.007. N.H. Van Dijk, High-energy X-ray diffraction study on the temperature-dependent
[3] M. Kazeminezhad, A. Karimi Taheri, The effect of controlled cooling after hot mechanical stability of retained austenite in low-alloyed TRIP steels, Acta Mater. 60
rolling on the mechanical properties of a commercial high carbon steel wire rod, (2012) 565–577, http://dx.doi.org/10.1016/j.actamat.2011.10.019.
Mater. Des. 24 (2003) 415–421, http://dx.doi.org/10.1016/S0261-3069(03) [33] P. Błoński, A. Kiejna, Structural, electronic, and magnetic properties of bcc iron
00095-5. surfaces, Surf. Sci. 601 (2007) 123–133, http://dx.doi.org/10.1016/j.susc.2006.09.
[4] M. Dollar, I.M. Bernstein, A.W. Thompson, Influence of deformation substructure on 013.
flow and fracture of fully pearlitic steel, Acta Metall. 36 (1988) 311–320, http://dx. [34] V. Mikulich, C. Blochwitz, W. Skrotzki, W. Tirschler, Influence of texture on the
doi.org/10.1016/0001-6160(88)90008-9. short fatigue crack growth in austenitic stainless steel, Mater. Sci. 42 (2006) 84–94.
[5] B.G. Bock, Guidelines to Best Practices For Heavy Haul Railway Operations: Wheel [35] G.I. Taylor, Plastic Strain in Metals, 1938.

154
M. Masoumi et al. Materials Science & Engineering A 722 (2018) 147–155

[36] S.I. Wright, M.M. Nowell, S.P. Lindeman, P.P. Camus, M. De Graef, M.A. Jackson, (2013) 3809–3820, http://dx.doi.org/10.1080/14786435.2013.798048.
Ultramicroscopy Introduction and comparison of new EBSD post-processing meth- [41] Z.Q. Wang, I.J. Beyerlein, R. LeSar, Slip band formation and mobile dislocation
odologies, Ultra. J. 159 (2015) 81–94. density generation in high rate deformation of single fcc crystals, Philos. Mag. 88
[37] R.I. Babicheva, S.V. Dmitriev, D.V. Bachurin, N. Srikanth, Y. Zhang, S.W. Kok, (2008) 1321–1343, http://dx.doi.org/10.1080/14786430802129833.
K. Zhou, Effect of grain boundary segregation of Co or Ti on cyclic deformation of [42] J.R. Low, A.M. Turkalo, Slip band structure and dislocation multiplication in si-
aluminium bi-crystals, Int. J. Fatigue 102 (2017) 270–281, http://dx.doi.org/10. licon-iron crystal, Acta Metall. 10 (1962) 215–228.
1016/j.ijfatigue.2017.01.038. [43] X. Chen, Z. Liu, P. Xia, A. Ning, S. Zeng, Transition of crack propagation from a
[38] A. Kedharnath, A.S. Panwar, R. Kapoor, Molecular dynamics simulation of the in- transgranular to an intergranular path in an overaged Al-Zn-Mg-Cu alloy during
teraction of a nano-scale crack with grain boundaries in α-Fe, Comput. Mater. Sci. cyclic loading, Met. Mater. Int. 19 (2013) 197–203, http://dx.doi.org/10.1007/
137 (2017) 85–99. s12540-013-2009-y.
[39] L.L. Li, Z.J. Zhang, P. Zhang, J.B. Yang, Z.F. Zhang, Distinct fatigue cracking modes [44] M. Masoumi, L.P.M. Santos, I.N. Bastos, S.S.M. Tavares, M.J.G. da Silva, H.F.G. de
of grain boundaries with coplanar slip systems, Acta Mater. 120 (2016) 120–129, Abreu, Texture and grain boundary study in high strength Fe-18Ni-Co steel related
http://dx.doi.org/10.1016/j.actamat.2016.06.032. to hydrogen embrittlement, Mater. Des. 91 (2016) 90–97, http://dx.doi.org/10.
[40] L.M. Brown, Cracks and extrusions caused by persistent slip bands, Philos. Mag. 93 1016/j.matdes.2015.11.093.

155

You might also like