You are on page 1of 22

Rail defects: an overview

D. F. CANNON
1
, K. - O. EDEL
2
, S. L. GRASSI E
3
and K. SAWLEY
4
1
Derbyshire, UK,
2
Fachhochschule Brandenburg, Germany,
3
Wedemark, Germany,
4
National Research Council, Centre for Surface Transportation
Technology, Ottowa, Ontario, KIV1S2, Canada
Received in final form 2 June 2003
ABSTRACT For about 150 years, the steel rail has been at the very heart of the worlds railway systems.
The rail works in a harsh environment and, as part of the track structure, it has little redun-
dancy; thus, its failure may lead to catastrophic derailment of vehicles, the consequences
of which can include death, injury, costs and loss of public confidence. These can have
devastating and long-lasting effects on the industry. Despite the advances being made in
railway permanent way engineering, inspection and rail-making technology, continually
increasing service demands have resulted in rail failure continuing to be a substantial eco-
nomic burden and a threat to the safe operation of virtually every railway in the world.
This paper presents an overview of rail defects and their consequences from the earliest
days of railways to the present day.
Keywords defects; grinding; inspection; rail; RCF
I NT RODUCT I ON
For about 150 years, the steel rail has been at the very
heart of the worlds railway systems. Its main functions
are to transmit wheel forces to the track bed and, together
with the tread and flange of the wheel, to guide vehicles.
The environment in which the rail works is harsh and the
forces endured by it are complex and variable. Wheelrail
contact conditions can result in severe wear, the environ-
ment may lead to corrosion and the rail may even be sub-
jected to mechanical and thermal abuse during installation
and track maintenance operations. The rail must also be
machinable, weldable and, because of the large tonnages
continually consumed, affordable. By and large the steel
rail has given good service; however, like many metallic
components that are subjected to cyclic (repeated) load-
ing, the rail is susceptible to metal fatigue and this can
lead to its partial or complete failure. As a structural unit
within the track construction, the rail has very little re-
dundancy; consequently, its failure demands immediate
rectification. At worst it may result in catastrophic derail-
ment of vehicles. The consequences of such derailments
in terms of death, injury, cost and public confidence can
be devastating and long lasting.
Fatigue failure develops in three basic phases; first a fa-
tigue crack initiates, it then grows in size, and, in the ab-
sence of control, the rail finally breaks (Fig. 1). The first
Correspondence: Dr. K. Sawley. E-mail: Kevin sawley@ttci.aar.com
two phases of fatigue failure require the accumulation of
loading cycles over a period of time. It is this period of
crack growth that is used by railway engineers to control
and contain the problem within limits that are considered
to be reasonable. This control process involves regular
rail inspection and the implementation of prescribed ac-
tions whenfatigue cracks andother defects are found. This
control process is often known as rail defect management
(RDM).
Despite many advances in railway permanent way en-
gineering, inspection and rail-making technology, con-
tinually increasing service demands have resulted in
rail fatigue failure remaining a substantial economic
burden and a threat to the safe operation of virtu-
ally every railway of the world. The universal nature
of the problem and its cost, probably around C
=
2 bil-
lion per year in the European Union alone, is indi-
cated by the decision of the International Union of
Railways (UIC) to make Rail Defect Management its
first World Joint Research Project.
1
This paper presents
an overview of rail defects and their consequences
from the earliest days of railways to the present day.
RAI L S, ST RE SSE S, RAI L FAI L URE AND COST S
The early days
The first metal rails, for example, as used in English
coalmines and tramways in the 18th century, were made
of cast iron. This material was brittle and was unable to
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865887 865
866 D. F. CANNON et al.
Fig. 1 Tache ovale or kidney rail failure.
redistribute loads through plastic deformation. With
hindsight, it is not surprising to find that rail failures were
common. Richard Trevithick experienced such rail fail-
ures with his first steam engine in the iron works of South
Wales and spectacularly, in 1808, a rail fractured during
the public demonstration in London of his locomotive
Catch me who can.
2
The rail failure caused the locomo-
tive to derail and, at least for the time being, presentations
of the new technology were ended.
3
Wear was also a problem. For example, it is believed that
George Stephenson increased the Killingworth colliery
gauge of 4 feet 8 inches by a 1/2 inch to reduce friction
between wheel flange and the rail, thus creating the stan-
dard track gauge of today (1435 mm).
4
In the interests
of reducing wear, the operating speed of the first German
railway, between N urnberg and F urth, was reduced from
the demonstration speed of 40 kmh
1
to 24 kmh
1
.
4
The
problem of wear is also clearly shown by the continuous
demand to increase the strength of rails.
Longitudinal woodenbeams cappedwithironstrips were
used as a cheap alternative to the iron rail in some early
19th century rail systems, for example, between Leipzig
and Wurzen in Germany and widely in North America.
This composite rail might almost be considered as the first
application of rail head hardening, nowcommonly used in
rail manufacture. The iron strips were about 25 mm thick
and they tried to bend as a result of deformation induced
by wheelrail contact stresses. The early nail fastenings,
later replaced by bolts, were unable to resist the bending
forces and the bent strips could become detached.
5
These
were known as snakeheads in North America, where they
caused frequent damage to equipment and passengers.
6
The advent and development of the steel rail
The steel rail became a practical proposition by the mid-
19th century. Bessemer, open hearth and other steel-
making processes made relatively high tonnage produc-
tion of steel possible, and by the turn of the century 15 m
Fig. 2 Wheelrail forces (taken from Ref. [14]).
lengths of rail were commonly rolled. In North America
early attempts at a composite, cheap rail (iron web and
base and steel head) soon gave way to adoption of the
complete steel rail. With the exception of the early bull
head rail, which continued to find favour in UK until the
mid 20th century, the flat-bottom or Tee-section rail, in
various sizes, has been almost universal for over a century
(see Fig. 2).
The performance of the steel rail, in particular its
strength and ductility, is much superior to the cast iron
rail and its advent opened the door to the rapid develop-
ment of railways throughout the world. The concept of a
steel rail supported by transverse beams, knownas sleepers
or cross-ties, set in stone ballast, remains the principle of
classic track design today even though ballastless track and
other new track forms are being used and developed.
Most of todays rail steels have basic carbon/manganese
compositions with pearlitic microstructures possibly with
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 867
Table 1 Chemical composition of Grade 260 Rail Steel
7
Chemical element (Mass %) ppm (max)
C Si Mn P (max) S Cr Al (max) V (max) N (max) O H
0.60/0.82 0.13/0.60 0.65/1.25 0.030 0.008/0.030 <0.15 0.004 0.030 0.008 20 3.0
Table 2 Mechanical properties of Grade 260 Rail Steel
7
Minimum ultimate Minimum Center line running
tensile strength (MPa) elongation (%) surface hardness (HBW)
880 10 260/300
some ferrite. The draft European Rail Standard grade 260
steel is typical of the steels used in many of the worlds
rail systems.
7
Its composition and basic material prop-
erties are shown in Tables 1 and 2. The microstructure
of this steel is described as pearlitic with possibly some
limited grain boundary ferrite. No martensite, bainite or
grain-boundary cementite is permitted. The draft Euro-
pean Standard also provides some minimumrequirements
for other properties that have a bearing on rail failure. For
grade 260 steel these are as follows.
Fracture toughness. The mean minimum toughness (K
Ic
)
should be 29 MPa m
1/2
at 20

C. The fracture tough-
ness of rail steels has a temperature and loading rate de-
pendency, and in the latter case the dependency may be
weak or strong.
812
Fatigue crack growth rates. The maximum fatigue crack
growth rates, measured in the head at room temperature
and at a ratio of minimumto maximumload of 0.5, should
be 17 m Gc
1
and 55 m Gc
1
at K values of 10 and
13.5 Mpa m
1/2
respectively.
Fatigue strength. The minimum fatigue resistance of rail
head material, determined in fully reversed axial loading
of small cylindrical test pieces, should be 0.000135 total
strain amplitude of 5 million cycles.
Residual stresses. The maximum longitudinal, tensile
residual stress in the rail foot, formed during manufac-
ture, should be less than 250 MPa.
Other than hardness, there are no requirements for wear
resistance. These properties indicate that todays typical
rail steel is strong and resistant to wear but its ductility is
limitedandat most operating temperatures it will fracture,
in the presence of a sharp tipped discontinuity, such as a
fatigue crack, in a brittle cleavage mode.
The draft European Rail Standard provides require-
ments for a number of other rail steels including some
with higher wear resistance achieved either by alloying
and/or heat treatment.
To meet todays stringent requirements, steel is normally
made by the basic oxygen process with vacuum degassing
and it is continuously cast. Electric-arc steel making and
secondary ladle arc refining may also be used. Great care
is taken to minimize the occurrence of atomic hydrogen
in the hot bloom, as this may lead to the formation of
shatter cracks in the rail head and the later formation of
the kidney rail failures (see Fig. 1). Slow bloom cool-
ing and isothermal treatment processes are also used to
ensure low hydrogen content. Rail rolling processes (e.g.,
the Bartscherer method) have beendeveloped to minimize
rolling flaws, and two-stage roller straightening is used to
maximize straightness and flatness, while minimizing lon-
gitudinal residual stresses.
For the future, bainitic steels showpromise of greater re-
sistance to rolling contact fatigue (RCF) damage, and im-
provements in joining processes, such as alumino-thermic
welds, are likely.
13
The basic rail section looks set to stay,
withlinear masses of around6070kgm
1
beingsufficient
for most applications. However, there may be more atten-
tion to the definition and accuracy of dimensions control-
ling the transverse shape of the rail running surface as this
affects the magnitude of contact stresses in the early days
of rail use, and hence RCF.
Rail stresses
There are many stresses that operate in a rail and can
influence rail defects and rail failure. Bending and shear
stresses arise principally fromthe gross vehicle load. Max-
imum static axle loads in Europe range from about 21 to
25 t but in the USA they routinely reach almost 30 t and
many coal trains running out of the Powder River Basin
have axle loads of about 32.4 t. In Australia axle loads of
about 37 t have been reported on iron-ore vehicles. All
these axle loads are nominal values, assuming that vehi-
cles are uniformly loaded. This need not be the case. Dy-
namic effects can significantly increase these static loads.
Conversely, if dynamic effects can be reduced, and loads
distributed more evenly, greater static loads can be car-
ried. As well as bending and shear stresses, the rail is also
subjected to contact stresses, thermal stresses and residual
stresses.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
868 D. F. CANNON et al.
Fig. 3 Head bending stress variation with wheel position (taken
from Ref. [14]).
Bending and shear stresses
As the wheel passes over a point in the rail, the forces it
applies produce vertical and lateral bending stresses and
shear stresses in the rail (Fig. 2).
14
While the lateral bend-
ing stresses contribute to rail failure, the vertical bending
stresses dominate. These are predominantly compressive
in the rail head and tensile in the rail base, although rail
uplift on either side of the wheel load position leads to a
reversal of stress (Fig. 3). By design the rail shape is such
that rail head wear has little effect on stresses in the rail
base. This is because, although wear reduces the rail sec-
ond moment of area, it also moves the rail neutral axis
down in the rail. In contrast, wear does lead to a signifi-
cant increase in the mainly compressive head stresses. The
shear stresses produced in the rail are the main cause of
failures at boltholes machined in the rail web to join rails.
(On most railways, continuously welded rail has been in-
troduced to reduce numbers of boltholes.)
Analytical methods to calculate bending and shear stress
generally assume that the rail is supportedona continuous,
elastic foundation of constant modulus; thus the effect of
discrete supports, i.e., sleepers, is not taken directly into
account. (Sleeper spacing has an indirect effect through
its effect on elastic foundation modulus.) This is a rea-
sonably good approximation. Finite-element methods are
also available to calculate stress.
Wheelrail contact stresses
Contact stresses between the wheel and rail are very high.
For example, the maximum normal stress can routinely
reach 1500 MPa compressive, while stresses in excess of
4000 MPa can arise from poorly conforming wheels and
rails. These stresses can usually be reasonably predicted
from Hertzian analysis, although more accurate analysis
methods are available that may indicate multipoint contact
especially in curving conditions (see Fig. 4).
15
Forces that
contribute to contact stresses are wheel load, and forces
Fig. 4 Normal contact pressure distributions, (a) Hertzian (b)
non-Hertzian (taken from Ref. [15]).
in the plane of the wheelrail contact area, which arise
from traction, braking and steering. The steering forces,
and the way they are generated, are complex, and have
components along and across the rail as well as a spin mo-
ment about an axis normal to the contact surface. These
in-plane forces have a significant influence on RCF.
Thermal stresses
To eliminate problems with bolted rail joints, most mod-
ern railways use rails that are welded together, known as
continuously welded rail. This has produced its own prob-
lems. The rail/sleeper system behaves very much like a
long slender column, and in compression it can buckle,
giving a high risk of derailment. For this reason, rails are
welded under conditions to simulate nominally high am-
bient temperatures, so that they are in tension for most of
the year. Compression only occurs in hot summer months
when the constrained rails would, if they could, thermally
expand. This method of installing rails reduces the risk of
buckling, but increases the risk of sudden brittle rail fail-
ure, especially in winter months when temperatures are
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 869
low and thermally induced tensile forces are highest. This
increases the need for non-destructive inspection to de-
tect defects. Thermally induced longitudinal stresses can
also occur in rails joined by bolts and fishplates (joint bars)
if joints are not well maintained.
Residual stresses
Residual stresses are introduced in the rail during manu-
facturing and straightening processes. These stresses are
subsequently modified by plastic deformations occurring
in the wheelrail contact zone. Residual stresses tend to
be tensile in the centre of the rail head, where they can
contribute to defect initiation and growth. The measure-
ment of residual stresses is not easy. Traditional destruc-
tive methods (e.g., strain gauge and sectioning, Meier
rods etc.) are relatively coarse and unsuitable when steep
stress gradients and multiaxial stresses are present. Non-
destructive methods, suchas ultrasonics andneutronbeam
diffraction have been used with mixed results.
16,17
Dynamic effects
Static forces and rail stresses are increased significantly by
track-top irregularities and discontinuities in the rails and
in the wheels. For example, joints and welds in rails can be
dipped when they are produced or dips may form due to
deformation and/or wear. These irregularities create high
dynamic forces, especially at high speeds. Similarly, high
dynamic forces can arise from wheels that have flat spots
or are out-of-round. Amajor cause of such irregularities is
wheel slide, caused when the wheel stops rotating because
Fig. 5 Surface spalling of the rail surface caused by wheelburn.
of braking demand that exceeds the available wheelrail
friction.
Dynamic forces can significantly reduce the critical size
of fatigue cracks. If they occur sufficiently frequently, they
can increase fatigue crack growth rates. Wheel-impact de-
tectors are therefore used in many countries. In North
America approximately 50 wheel-impact detectors were
in operation at the end of 2002. Approximately 10 other
systems able to measure vertical and lateral wheel forces
were expected to be operational to detect poorly perform-
ing bogies that cause high lateral forces. The Association
of American Railroads interchange rules condemn a wheel
if a force of about 400 kN is detected; for a typical-loaded
freight wagon this represents a dynamic force increment
of about 2.7.
Types of modern rail failure
Todays rail failures can be divided into three broad groups
as follows.
r
Those originating fromrail manufacturing defects a clas-
sic example of this is the tache ovale or kidney defect that
usually originates from a hydrogen shatter crack in the rail
head (Fig. 1).
r
Those originating from defects or damage caused by inap-
propriate handling, installation and use. For example, the
wheelburn defect (Fig. 5) is caused by spinning wheels.
r
Those caused by the exhaustion of the rail steels inher-
ent resistance to fatigue damage. Many forms of RCF are
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
870 D. F. CANNON et al.
Fig. 6 Surface gauge corner head checks.
Fig. 7 Transverse crack originating at head checks.
within this group, for example head checking (Figs 6 and
7) and squats (Fig. 8).
Rail failures caused by manufacturing defects have been,
and are being, addressed through improvements in steel-
and rail-making technology. On the other hand reduction
of rail failures caused by abuse lies in the hands of the
railways and railway service industries. Defects and rail
failures in these two groups have traditionally been the
major target of RDM(see the section Rail defect manage-
ment). Over the last 20 years or so, rail failures in the third
group have been increasing, especially failures involving
RCF. Defects and rail failures in the third group can only
be effectively reduced by improvements in material prop-
erties, changes in the design, build and maintenance of
the track and vehicles or changes in the conditions of
operation.
Rolling contact fatigue
The October 2000 derailment at Hatfield inUK, resulting
from head checking, provides an outstanding example of
the possible consequences of RCF. RCF is likely to be
a major future concern as business demands for higher
speed, higher axle loads, higher traffic density and higher
tractive forces increase.
Head checks, gauge-corner cracks and squats are all
names for surface-initiated RCF defects. They are caused
by a combination of high normal and tangential stresses
between the wheel and rail, which cause severe shearing of
the surface layer of the rail and either fatigue or exhaus-
tion of ductility of the material. The microscopic crack
produced propagates through the heavily deformed (and
orthotropic) surface layers of steel at a shallow angle to
the rail running surface (about 10

) until it reaches a depth


c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 871
Fig. 8 Squat defect with developed transverse crack.
where the steel retains its original isotropic properties. At
this stage the crack is a few millimetres deep into the rail
head. At this point the crack may simply lead to spalling of
material from the rail surface. However, for reasons still
not clearly understood, isolated cracks can turn down into
the rail, and, if not detected, cause the rail to break. These
events appear to be rare, but are highly dangerous since
RCF cracks tend to form almost continuously at a given
site. Fracture at one crack increases stress in the nearby
rail, increasing the risk of further breaks and disintegra-
tion of the rail.
Classical work by Bower and Johnson
18
and Bogdanski
et al.
19
has shown that early, shallow angle crack prop-
agation, is encouraged by fluid entrapment that leads to
crack pressurization and to reduced crack face friction that
allows relative shear of the crack faces. The direction of
growth of the crack beneath the rail surface and the di-
rection of the crack mouth on the rail surface are both a
guide to the predominant direction of traction causing the
crack.
RCF initiation is not normally associated with any spe-
cific metallurgical, mechanical or thermal fault; it is simply
a result of the steels inability to sustain the imposed oper-
ating conditions. The problem is known to occur in most
of the rail-steel types in common use today. Thousands of
fine surface head checks can form on the high rail of some
curves (Fig. 6) and deep transverse head cracks may extend
from a few of them (Fig. 7). The squat defect is a similar
Fig. 9 Shadowing of ultrasonic sound by head checks and squats.
formof fatigue damage that tends to occur more randomly
on very shallow curves and tangent track (Fig. 8).
These forms of surface-initiated RCF pose special in-
spection problems. For both head checks and squats the
development of a downward-turning fatigue crack leads to
final rail failure. Unfortunately, the earlier shallow-crack
development phase can shadow this from conventional
ultrasonic examination (Fig. 9). Better ultrasonic probe
arrangements and the use of eddy current technology are
under investigation as possible solutions.
The problem can be contained by grinding the rail run-
ning surfaces to remove fatigue damaged material and/or
to impose an improved rail head transverse profile, that is,
one that reduces contact stress or improves steering (see
the section Rail grinding).
RCF can also be limited by reducing wheelrail tractions
in curves through lubrication, or friction management
more generally. Although water entrapment has probably
the biggest influence on shallow crack growth through
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
872 D. F. CANNON et al.
fluid entrapment, where there is an absence of water (in
underground railways and in otherwise dry environments)
grease-based lubrication may also influence the growth of
surface-initiated cracks through fluid entrapment. Using
lubricationtoreduce frictiononthe railheadis alsofraught
with difficulty, because coefficients of friction greater than
0.25 are commonly required for reliable traction and
braking.
In the late 1990s RCF defects accounted for about 60%
of defects found by East Japan Railways, while in France
(SNCF) and UK (Railtrack) the figures were about 25
and 15%, respectively.
20
At that time, in UK, the RCF
defects were classified as squats. The RCF problem was
experienced in the 1960s on Japans high speed Tokaido
Shinkansen line and by the early 1970s it had already
become a significant reason for rail replacement on the
British Railways West Coast Main Line that had been up-
graded and electrified in the 1960s.
Heavy haul aspects
The experience of so-called heavy-haul railways is dif-
ferent from that of passenger and mixed-passenger and
freight railways, such as those in continental Europe,
Japan and UK. Heavy-haul railways, which are character-
ized by high axle-load traffic, uniform rolling stock, long
trains and high annual tonnages of traffic, tend not to use
the same rail failure reporting methodologies used by pas-
senger railways, and sometimes use different terminolo-
gies for defects (see the section Rail defect management).
Heavy axle loads can lead to defects not typically seen in
passenger railways. The principal one of these is the trans-
verse defect (TD) that initiates underneath the high-rail
gauge corner, typically from inclusions in the rail steel.
Changes in types of rail defect that have occurred from
heavy-haul operations is illustrated in Table 3 by a com-
parisonof the types of defects foundinUSrailways in1948
(before the inception of heavy-haul operations) and in
19901995, when heavy-haul operation was widespread.
In 1990/1995, TDs comprised the largest single category
of rail defects, whereas in 1948 there is neither a specific
category of transverse defect nor a category that describes
defects that are substantially similar to the TD. It can
reasonably be concluded that this type of rail failure has
blossomed as a result of heavy-haul operations.
Apart from this striking difference in behaviour, there
are great similarities between the types of defects in 1948
and in the early 1990s. In particular, it is clear that the
joint between two rails is a major cause of defects. In
1948, the single greatest category of defects (Web defect
at joint) is presumably a description of bolthole failures at
rail joints, and a further 15.9% of defects are associated
withthe rail webandcouldalsobe a descriptionof bolthole
failures. In 1990/1995, boltholes and defective rail welds
Table 3 Comparison of 1948 US defect types with 1990/1995
defect types
Defects reported in 1948
a
Defects for 1990/1995
Defect type % Defect type %
Web defect at joint 36.2 Transverse defect 26.0
Compound fissure/detail 18.9 Defective thermic weld 17.4
fracture
Other web defects 15.9 Bolt-hole defect 16.1
Broken rail 8.1 Vertically split head 9.6
Vertically split head 6.2 Defective flash weld 7.9
Horizontally split head 5.7 Head/web separation 6.1
Other head defects 5.7 Horizontally split head 5.6
Base defects 3.2 Engine burn defect 3.0
Split web 3.0
Others 5.2
Total 100
a
From rail rolled in the previous ten years.
accounted for about 42% of total rail defects. Improve-
ment of alumino-thermic welds would contribute greatly
to the reduction of rail defects, in North America and
worldwide.
Rail failures and accidents
Statistics of rails found damaged, cracked or broken are
available from many railways; for example, data from the
mid to late 1990s fromthe UK, France, Netherlands, Ger-
many, Japan, South Africa and India can be found in Ref.
[20]. Some RDM cost information is also available, but
in general, there is little information concerning the costs
incurred when rail breaks cause derailments or other acci-
dents (the USA is a notable exception to this generality).
This lack of data may arise froma reluctance of railways to
ascribe specific causes to accidents. In some circumstances
it may be difficult to prove that a rail break preceded an
accident and was either a partial or principal cause of it.
Loss of part or all of the rail running surface greatly raises
the risk of derailment. Bolthole cracks at rail joints and
any other defect causing cracks to propagate in the rails
longitudinal directionare therefore a major hazard. All rail
defects that can occur in clusters, froma fewmillimetres to
over a few metres, and that may cause multiple transverse
rail fracture, are also a major threat. These defects include
the kidney fracture (Fig. 1), various forms of RCF damage
(Figs 68) and wheelburn (Fig. 5). These multiple defects
can lead to complete disintegration of the rail over many
metres.
Rail defects that lead to isolated transverse fractures are
less likely to cause train derailment. Even in continu-
ously weldedrail, subject tohighthermally inducedforces,
transverse fractures normally result in gaps of no more
than about 30 mm during the winter and less at other
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 873
times. (At very lowtemperatures and/or with poor sleeper
and/or rail fastening conditions larger gaps can occur.)
However, transverse fracture causes a loss of load path,
and a step is presented to the oncoming wheel that leads to
deformationof the running-onend. Rail ends may eventu-
ally become sufficiently mis-aligned to cause derailment.
Poor ballast, sleepers and fastenings, especially in curved
track, exacerbate the problem.
German data from the turn of the 19th to the 20th cen-
tury indicated an accident/rail break ratio of about 0.0007
and 1965 data reported for various European railways in-
dicated ratios ranging from 0.0039 to 0.00046.
21,22
How-
ever, it should be noted that the railway reporting the low-
est ratio also reported the largest number of rail breaks,
39 000, compared to the next largest of 7377. In 1972
a study in the United States indicated an overall acci-
dent/rail failure ratio of 0.0044 but these data probably
included defective and broken rails.
23
Data from 1989 for
a large US railway with about 26 000 miles of track indi-
cated an accident/rail break ratio of about 0.0034.
24
Taking the typical European data of 1965 gives an acci-
dent/rail break ratiorange of 0.00070.0039, withinwhich
the more recent US railway data fall. Perhaps surprisingly,
the somewhat limited data suggest that the ratio of acci-
dents to rail breaks has remained fairly constant with time
and from one railway to another.
This ratio is also a useful indication of the relative sever-
ity of breaks arising from different causes. For example,
data reported for five European railways show high acci-
dent/rail break ratios for horizontal web and wheelburn
defects (0.04 and 0.033 respectively), while the ratios for
transverse fractures and alumino-thermic weld fractures
are relatively low (0.0014 and 0.00043, respectively).
22
It is assumed that the data above are derived from plain
rail breaks and welds (for example, the UIC Catalogue
of Rail Defects only applies to these cases). However, al-
though switches and crossings comprise but a small part of
the overall length of railway track, they have been associ-
ated with substantial number of rail removals. Dearden re-
portedfor BritishRailways that ratios of rail removedfrom
plain track and switches and crossings were 2.5 and 1.8 for
the years 1952/1954 and 1961/1963, respectively.
25,26
Rail life and causes of removal
The average life of rail in track depends on many factors
including rail quality, vehicletrack interaction, renewal
and maintenance policies, and even local economics in-
cluding relative capital and labour costs. Table 4 gives the
estimated life of rail in heavy haul straight and curved
track. These figures are from a survey of major North
American railways conducted in 2002. The main factor
influencing rail life varies between railways, but another
survey of North American railways in 2000 indicated that
Table 4 Estimated rail lives in heavy-haul track
Estimated rail
Curve radius life (MGT traffic)
Straight 1073
1164 865
699 699
499 605
388 510
318 449
269 408
233 380
206 358
184 334
<175 251
rail wear was the main cause of rail replacement (Table 5).
The figures show, however, that a significant amount of
rail is purchased to replace rail with defects, including sur-
face damage.
Rail breaks, RCF and side wear in curves limit rail life.
Germandata from1937 indicate that there were then20%
more breaks per unit length of track in curves than in
tangent track.
27
This figure is likely to be higher now as a
result of the increasing occurrence of RCF defects (head
checks and gauge corner cracks).
Costs of rail failure
The costs of rail failures include the following.
r
Inspection for example visual and ultrasonic. These costs
depend on the frequency of inspection; this may at least
vary from six times a year in main lines to once every
9 years in lightly used lines.
r
Train delay these costs especially occur when penalties
are payable by the track authority and its contractors to
train operating companies.
r
Remedial treatments for example rail replacement, weld
repair.
r
Pre-emptive treatments for example rail grinding.
r
Derailments.
r
Loss of business confidence and customer support.
Some of these costs, at least in specific cases, can be
established with some certainty, for example inspection
costs, but others cannot. Derailments are relatively rare
in high-speed passenger networks but they may be more
frequent in low-speed freight systems. However, the cost
of a high-speed passenger train derailment is likely to be
much greater, indeed it may overwhelm all other costs.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
874 D. F. CANNON et al.
Table 5 Reasons for rail purchase on North American heavy-haul railways
Percentage of rail purchased annually
Reasons for rail purchase Railway A Railway B Railway C Railway D
Rail at its wear limit 25.3 85 58.8 85
Rail exceeding defect limits 6.5 29.5 5
Damaged rail (e.g., derailments) 1.7 7.8 0
Rail with surface damage 21.2 15
New lines and upgrades 45.3 3.9 10

There is uncertainty in the reliability of rail failure


statistics. Moreover, defect descriptions and classifications
differ; a situation that led the UIC to publish an Inter-
national Cross-Reference of Rail Defects (see subsection
Reporting and classification of rail failures and Ref. [32]).
Traffic and track conditions vary and even the accuracy of
reporting defects may be questionable; thus caution needs
to be exercised when collating and comparing rail failure
statistics.
Despite the difficulties outlined above it is clearly valu-
able to try and get some feel for the magnitude and cost
of the rail failure problem. In Ref. [20] a global figure of
about 0.1 defective and broken rails per track kilometre
per year was suggested with a ratio of about 10 to 20 to
one between defective and broken rails. These global fig-
ures were derived fromrail systems throughout the world;
nonetheless, they provide some indication of the scale of
the rail failure problem.
Using the above global failure rates and cost data from
various railways it was estimated in Ref. [20] that the over-
all cost of rail failure, excluding pre-emptive treatments
and derailments, was of the order of C
=
7501700 per kilo-
metre per year. Total track length in the European Union
is about 0.5 million kilometres; thus the total cost, assum-
ing that rails are inspected annually, is around C
=
375850
million per year. As these figures exclude rail grinding and
derailment costs they are clearly on the low side. Else-
where it was estimated that RCF defects alone cost rail-
ways within the European Union around C
=
300 million
per year and, as these defects probably account for about
15% of the total, the emerging cost of all defects is about
C
=
2 billion per year.
28
Assuming annual vehicle ultrasonic inspections followed
by manual verificationof detecteddefects, inspectioncosts
are estimated at about C
=
70 million per year for a 0.5 mil-
lion kilometre system.
20
Thus costs arising from the pres-
ence of defects range from about 4 to 27 times that
amount. Reduction of defect numbers reduces these costs
even if inspection costs remain unchanged.
The above figures can only be taken as a very rough
estimate of the cost of rail failures, but they indicate the
magnitude of the problem.
These estimates above do not include derailment costs
that can figure more prominently in heavy-haul systems.
In the USA, for example, approximately 290 derailments
reported by the railways to the Federal Railroad Admin-
istration in 2001 were ascribed to broken rails. The es-
timated costs of these derailments were about C
=
97 mil-
lion. The Association of American Railroads uses approx-
imately C
=
350 per train delay hour as a cost for freight
train delays. This can be compared to around C
=
250 000
per hour at a key passenger junction in Europe. There is
thus considerable safety and economic pressure to reduce
numbers of broken and defective rails.
RAI L DE F E CT MANAGE ME NT
Costs of rail failure are high; therefore, substantial efforts
are made to control it. This control process is sometimes
known as RDM and features of it are briefly described
below.
Regulations
Regulations affecting RDM can be found in laws, de-
crees, directives, rules and instructions issued by many
different bodies. Only a few examples are given below for
illustration.
Governments may impose regulations affecting RDM at
various levels. At the highest level this usually takes the
form of broad legislation. For example, in the UK the
Transport Act requires railways to operate with due re-
gards to efficiency, economy and safety. Legislation of
this kind sets expectations but it firmly places the practi-
cal responsibility for safe operation with other regulatory
bodies and/or the railway businesses.
Regulatory/advisory bodies can take many forms. In the
United States the Department of Transportation Federal
RailroadAdministration(FRA) is one suchbody. The FRA
sets basic track safety and inspection requirements in their
Track Standards Handbook and rail defects resulting in
train derailment are reportable to the FRA, along with
cost/casualty data. In the UK the Health and Safety Exec-
utive enforces relevant legislation and plays a major role
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 875
in the investigation of serious accidents. In the Nether-
lands the Ministry of Transport prescribes minimum
ultrasonic rail inspection frequencies and other safety
requirements.
High-level control by regulatory bodies has the advan-
tage of freedom from immediate commercial interests;
however, it may be viewed as cumbersome if checks and
controls are detailed. For example: responsibility is remote
from operations; high-level control may reduce flexibil-
ity and inhibit performance-based response; time inertia
may inhibit effective responses to developing rail-defect
trends.
For many railways, high-level control imposes funda-
mental obligations that are met through detailed in-house
regulations, handbooks, manuals, specifications etc. Rail-
ways may also join together, perhaps through an associ-
ation, to create regulations and specifications. At an in-
ternational level the UIC has produced documents that
can be applied to RDM; these include the Catalogue of
Rail Defects that describes and classifies rail defects.
29
It
also contains recommended minimum actions to be taken
on finding a defect. The actions given in Ref. [29] may
be considered vague and of limited practical value; no
doubt reflecting the difficulty of providing universal ad-
vice, where targeted requirements, taking note of actual
track and traffic conditions, are necessary.
Objectives of rail defect management
Two essential objectives of RDM are: to detect and rectify
rail defects before they cause rail breaks and to reduce and
eliminate rail defects.
In general RDM strategies have tended to focus on the
first objective, and, until the last 10 or 20 years, this has
probably been reasonable. Many rail defects and rail fail-
ures result from manufacturing problems and use and
abuse of rails in track, and many of these problems have
been, and are being, addressed. For example, improved
manufacturing technology has reduced the presence of
hydrogen and hard brittle sub-surface inclusions such as
aluminates and silicates, reducing fatigue crack initiation
sites. The move to continuously welded track has elimi-
nated most bolted joints and the risk of fatigue cracks at
joint boltholes. These and other measures should con-
tinue to reduce rail defects and rail failures. However,
this scenario has been changing with the recognition and
widespread emergence of surface initiated RCF defects
such as squats and head checks.
In recent years a number of railways, principally but not
always passenger railways, have adopted the goal of zero
rail breaks. Today, this may not be practically possible,
but it is a goal that drives reductions in breaks and the
development of RDMsystems andwheelrail technology.
Reporting and classification of rail failures
The system used to report rail failures and analyse rail-
failure statistics is an essential part of RDM. A rail is de-
fined as failed when it is broken, cracked or otherwise
damaged and can no longer fulfil its design function. This
definition also applies to failed welded rail joints and weld
repairs.
Broken rail is deemed to occur when it is completely sep-
arated into two or more parts. A rail may also be called
broken when part of the running surface is lost. The cri-
teria may vary from one system to another but normally
involves the specification of a minimum length and depth
of the lost piece. Similarly rail with part of the foot missing
may be said to be broken. Broken rails require immediate
action to ensure safe operation until the rail is removed in
the shortest possible time.
Cracked rail is the partial or incomplete separation of the
rail.
Damaged rail is deemed to occur when the serviceability
of the rail is impaired by changes to its design geometry
and/or material properties or when such changes may lead
to the formation of cracks or breaks. Normal wear of the
rail is not usually treated as a failure.
Defect classification
When combined with statistical analysis and material test-
ing, systematic classification of rail defects allows conclu-
sions to be drawn about: the essential causes of failures and
conditions for their occurrence; the introduction of mea-
sures to combat failures; the effectiveness of the measures
introduced and allowable defect sizes.
Classification of a rail failure is most likely to be under-
taken by track engineering or rail inspection staff and the
data created is normally stored and manipulated by com-
puters. Since laboratory facilities are not always available
for failure analysis, classification needs to be possible from
defect geometry alone. (Some rail businesses subject all or
samples of failures to laboratory examination since some
failures cannot otherwise be correctly identified.)
Many railways have used the UIC Catalogue of Rail De-
fects as the basis for their classification systems. This cat-
alogue was first published in 1959 and the fourth edition
was published in 2002.
29
It contains classifications for bro-
ken, cracked and damaged rails and rail welds. Afour-digit
code systemis used. The first digit indicates either the po-
sition of the defect in the rail (1 at the rail end; 2 away
from rail end) or special types of defect (3 defects result-
ing from damage to the rail; 4 welds and re-surfacing
defects).
The second digit indicates the place in the rail section,
where the defect started or the welding method (if the first
digit is 4). The third digit indicates the crack geometry or
type of defect or its origin or cause. For example, if the
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
876 D. F. CANNON et al.
defect shown in Fig. 1 occurred away from the rail end
it would be classified as 211 (2 away from rail end; 1
rail head; 1 transverse; in this example there is no fourth
digit).
In 1961 the Soviet Railways and the Eastern Bloc Coun-
tries also produced a defect catalogue through the co-
operative organisation OSShD.
30
This used three digits
for coding defects; the first gave the type of defect and
its position in the rail cross-section, the second gave the
sub-type of defect and the supposed cause and the third
gave the position of the defect along the rail.
Neither of the above two systems indicate the stage of
the damage when action was triggered (rail removal or
repair). To remedy this the OSShD catalogue was revised
in 1985 to include a fourth-digit code to indicate general
damage (1), crack (2) or break (3).
31
The Sperry Rail Defect Catalogue is widely used in
North America and many other countries.
6
This manual
covers plain rail and welds but does not provide a classi-
fication system and this has led to defects being known
simply by their names or abbreviations.
The variety of classification systems used can lead to
problems whencomparing international experience. With
this in mind the UIC commissioned and published an in-
ternational cross reference of rail defects.
32
This man-
ual compares defect descriptions andclassificationsystems
used in a number of countries around the world.
Inspection and actions
Rail breaks are normally detected visually, for example by
inspectors or other track staff, by breaks in electrical track
signalling circuits that the rails form part of, and by train
operating staff (and even passengers) experiencing uneven
or unusual ride behaviour. Rail breaks are dealt with as
emergencies either by temporary or permanent repair or
replacement of the rail.
RDM aims to prevent rail failure by detecting defects by
inspection before the rail breaks. A measurement of de-
fect management efficiency is the ratio of broken rails to
detected defects. A high ratio means that many rails break
before defects are found. A low ratio means that inspec-
tion methods are effective in finding defects. The ratio
changes with defect type, but the US data from three rail-
ways indicate a ratio of broken to detected defects gener-
ally less than0.1, and oftenmuchless. Following detection
of a defect various actions may be described, ranging from
keeping the defect under observationtoimpositionof traf-
fic speed restrictions to repair or renewal immediately or
within 1 or 2 days of detection. Instant repair/renewal
(with no further traffic following detection) is relatively
rare in high-density passenger systems, but is more com-
mon in heavy-haul systems, where high wheelrail forces
may result in high defect growth rates. Prevailing regu-
lations and the legal liability environment are also fac-
tors in determining actions. For example, in the USA the
Federal Railroad Administration regulations require rails
withcertainidentified defects to be either removed or sup-
ported with emergency joint bars. Speed restrictions often
also need to be applied. The result is that, for both safety
and economic reasons, many heavy-haul railways verify
automatically detected defects by hand, and immediately
remove/repair rails with significant defects.
When rails containing defects are allowed to remain in
track, the actions following detection are designed to pro-
vide time for rail repair or replacement, while minimiz-
ing traffic interruptions and, critically, minimizing the risk
of the rail breaking. This is achieved by specifying a re-
inspection interval or a time limit for a rectification or rail
removal operation.
These actions are often termed minimum actions, and
they are the minimum that should be done. They do not
stop the track engineer taking more urgent or appropriate
action if, for example, the consequences of a broken rail
would be especially severe. Actions have largely evolved
over time and setting them depends a great deal on expe-
rience and knowledge gained over many years. The UIC
Handbook of Rail Defects describes some minimum ac-
tions, but it is usual for individual rail systems to set their
own, bearing in mind local conditions and statutory leg-
islation and the requirements of all regulatory bodies.
29
Currently, explicit or formal cost/risk analysis plays very
little part in the practical determination of actions, rail
inspection procedures and RDM.
The German Railways operate actions, which are quite
typical.
33,34
For example, detection of a transverse rail
head defect (e.g., Fig. 1) leads to actions that depend
upon: line speed; area of the defect for line speeds less
than 160 kmh
1
; the ultrasonically indicated height of the
defect for line speeds greater than 160 km h
1
; distance
between defects and line speed (in the case of multiple
defects).
Larger defects require immediate securing of line safety
and the removal of the rail; smaller defects may remain
in track until the next inspection when any evidence of
growth will require the rails removal. For this particular
type of defect areas between 20 and 55% of the rail head
area have been reported.
3537
The size of defect prompt-
ing rail removal varies from railway to railway.
Inspection intervals may be set in accordance with op-
erational conditions; including, for example, line speed,
track construction, traffic tonnage and axle load, derail-
ment risks and traffic type (including whether hazardous
materials are transported).
For example, German Railways specify inspection in-
tervals from 4 to 24 months.
34
Most of the large freight
railways in North America now use some form of risk
management to schedule ultrasonic rail inspections, based
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 877
on outputs from simple formulae. A 40 million gross
tonnes (MGT) per year freight line will be inspected two
or three times annually and very high tonnage lines, say
over 140 million tonnes per year, may be inspected every
30 days. Inspection intervals can be as frequent as every
7 days, for example, on Australian 37 tonnes axle load
lines.
Selection of inspection intervals is largely based on feed-
back of the end result, that is the number of defects found,
the number of rail breaks that occur, the number and
consequences of derailments caused and the perceived ac-
ceptability of the situation.
20
Major weaknesses of the ap-
proach are its failure to deal quickly and effectively with
emerging problems and the difficulty of assessing the out-
come, and relative merits, of alternative strategies. Im-
provements are likely to be best achieved by mathematical
modelling of the rail failure process and explicit cost/risk
analysis. As yet this approach plays only a small part, if
any, in determining inspection frequencies.
Good reporting and management control systems are es-
sential if actions are to be used safely and effectively. This
demands close and effective collaboration between the
controlling authority, inspection staff, and track mainte-
nance and renewal bodies. As part of this management sys-
temdefect records are increasingly logged into computer-
based local track maintenance programmes that are then
used to plan remedial work.
Reporting and use of rail failure statistics
Rail failure-report forms contain much associated and im-
mediately relevant data, for example about geographical
location, damage type, route, track and rail. Nonetheless
some essential data are not, or not exactly, recorded. For
example, the size of the fatigue crack leading to final rail
fracture would be helpful to the application of fracture
mechanics to RDM systems. Sources of faulty data and
data loss are the generally insufficient classification sys-
tem for rail failures and the fact that report forms must
often be submitted before faulty rails are removed and the
fracture face inspected. Also many report forms are likely
tobe completedby track engineers whoare not fatigue and
fracture specialists, and whose main concern is repairing
the track often under intense time pressure.
Periodic evaluation of rail-failure reports provides a
means of relating failures to steel type, manufacturer, date
of manufacture, traffic type, track loading and other pa-
rameters. This can help railways find the cause of rail fail-
ures, toimprove safety by the introductionof countermea-
sures and to prove the effectiveness of these measures. It
also aids optimisation of track maintenance. Bearing this
in mind more accurate information than the current qual-
itative state of rail and weld damage is required if rail fail-
ures are to be a measure of the effectiveness of diagnostic
and maintenance technologies as well as a measure of the
magnitude of rail removals and the avoidance of service
safety hazards.
RAI L I NSPE CT I ON
Inspection in production
Major improvements have been made to rail inspection
methods used in manufacture. Rails are examined visu-
ally for protrusions, hot marks, seams, scratches, rolled in
scale marks, cold marks and surface microstructural dam-
age. The underside of the rail foot is particularly sensitive
to defects that might initiate a fatigue crack and so this is
normally inspected automatically using, for example, eddy
current systems. Ultrasonic methods are used to check for
internal defects, especially in the rail head and web areas.
Only a limited area of the foot is normally inspected. Im-
proved production technology has led to major reductions
in manufacturing defects (see subsection The advent of
development of the steel rail).
Inspection in track
Since the 1960s, ultrasonic defect detection has been
widely used to detect surface breaking and internal de-
fects. The approach can also measure rail height and assess
corrosion of the foot. Unfortunately, small surface break-
ing defects cannot be reliably detected, and it is partly for
this reason that rail grinding techniques have been de-
veloped to remove certain types of RCF cracks (see Rail
grinding).
The principle of ultrasonic inspection is that a piezo-
electric transmitter injects a directed, high-frequency vi-
bration into the rail (at ultrasound frequencies) that may
be reflected or obscured by a defect in its path. A receiver
picks up signals from the transmitter and, depending on
signal strength and/or the relative positions of the re-
ceiver and transmitter it is possible to estimate the size,
and sometimes characteristics, of a defect. Ultrasonic de-
tection systems may be operated by pedestrians or more
complex systems may be vehicle based (Figs 10 & 11).
38
Ultrasonic methods can also be used to measure the de-
velopment of cracks in rails.
Vehicle based ultrasonic test systems operate typically
at speeds of 4070 km h
1
and new generation vehi-
cles at speeds up to 100 km h
1
. However, average
speeds are often much lower, especially in North Amer-
ica where immediate manual verification of defects and
possibly their repair or removal is frequently undertaken.
In such circumstances average speeds may be as low as
15 km h
1
.
The American Railway Engineering and Maintenance-
of-Way Association (AREMA) has set guidelines on the
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
878 D. F. CANNON et al.
Fig. 10 Pedestrian-operated test equipment (taken from Ref. [38]).
Fig. 11 Rail Testing Express DB Netz AG
(taken from Ref. [38]).
Fig. 12 Defect size and probability of detection using ultrasonic
inspection (taken from Ref. [39]).
probability of finding differently sized defects using ultra-
sonic inspection.
39
The guidelines for transverse defects
are shown in Fig. 12 (solid line). Also shown in Fig. 12 are
recent results from a study of the ability of modern ultra-
sonic inspectionvehicles to find transverse defects ranging
in size from 5 to 50% of the head of the rail. These re-
sults were found at the rail defect test site at the US FRAs
Transportation Technology Centre. The results showthat
the AREMA guidelines are both realistic and achievable.
In the latest ultrasonic systems, fully digital signal pro-
cessing is used to identify defects and provide additional
information to the operator; this should remove problems
arising from operator signal interpretation. Such soft-
ware has given a reduction in false alarms by using high
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 879
resolution recognizer logic, and sequence analysis, com-
bining signals from multiple transducers.
Other methods used to detect defects include the
following.
r
Visual inspection often carried out by track maintenance
staff and pedestrian operators of ultrasonic equipment.
r
Dye penetrant and magnetic particle methods used to
identify surface defects. Bothmethods are suitedtodetailed
manual inspection of rails.
r
Eddy current testing a non-contacting method of iden-
tifying surface breaking or near surface breaking defects
(depth may also be estimated).
r
Radiography used by some railways especially for the
examination of alumino-thermic welds that contain irreg-
ular, non-planar defects, and also for defects with orien-
tations unsuitable for ultrasonic inspection from the rail
top.
New inspection developments taking place include the
following.
r
Low-frequency eddy-current sensors to locate deeply
buried defects.
r
Neural network analysis of signals to improve defect de-
tection and identification.
r
Longitudinal guided waves, for example to potentially al-
low locomotives to scan the track ahead.
r
Laser generation and reception of ultrasonic waves to en-
able non-contacting inspection.
r
Improved ultrasonic probe combinations and arrange-
ments in vehicle-based systems.
r
Higher speed ultrasonic testing.
r
Non-destructive measurements of residual stresses and rail
neutral temperature.
RAI L GRI NDI NG
Grinding has been undertaken for many years to maintain
rail to increase rail life. Rail grinding objectives have in-
cluded the removal of corrugation (undulations on the rail
surface that increase dynamic forces), the removal of rail
surface damage (which also improves ultrasonic inspec-
tion), and rail re-profiling to improve vehicle steering. In
the last two decades increasing emphasis has been given
on grinding to remove cracks produced by RCF. Such
cracks form on almost all railways, from transit systems to
high-speed passenger railways and heavy-haul freight rail-
ways. While much theoretical and experimental work is
in progress to understand and combat RCF, most railways
see grinding as the only tool that is currently available to
control the development of small cracks into significant
defects. Rail grinding has developed as an essential part of
an effective treatment for RCF.
RCF and grinding
RCF of rails has increased in severity and extent over
the last 2030 years for several reasons including the
following.
r
Increasing traffic, increasing axle loads, and more uniform
trains particularly the combination of all three of these
in heavy-haul railways.
r
Lower wear rates resulting from improved lubrication and
harder rails. This has a twofold effect. First, if the installed
rail profile does not match well with the wheels passing
over, high stresses are produced until the rail wears to a
more suitable shape. Hence low wear rates imply longer
periods of high stress. Second, high wear rates can remove
embryonic cracks before they can propagate to significant
depths. Low wear rates prevent this.
r
The introduction of more powerful locomotives, particu-
larly those with traction control systems that are able to use
the available friction between wheel and rail without wheel
slip. The higher shear forces these locomotives apply to the
rail surface may lead to increased RCF.
r
In some rail systems there has been increasing deployment
of passenger vehicles designed more for high-speed stabil-
ity than good curving performance. There has accordingly
been a tendency to increase the tangential forces generated
in curves.
r
Wheel profiles, both circumferential and transverse, are
sometimes not well maintained. There has been a ten-
dency for this to happen where railway companies are no
longer vertically integrated, that is when there is no longer
single control of both the track and the vehicles running
over the track. Poorly maintained transverse profiles can
give deterioration in curving performance and high con-
tact stresses. Poor circumferential profiles (wheelflats and
out-of-roundness) give high dynamic loads on the rail.
Different types of RCF are described in subsection
Rolling contact fatigue, and have been reviewed else-
where.
40
There are basically two types: one initiating
on or very close to the surface and the other initiating
sub-surface. Surface-initiated defects exist, in different
forms, on almost all types of railway. Sub-surface initi-
ated defects, the most significant of which is the shell or
transverse defect, are associated primarily with heavy-haul
railways.
The value of grinding as a treatment for surface-initiated
RCF defects is threefold.
r
Grinding removes existing cracks, or curtails their length.
If grinding is undertaken frequently enough to maintain
the cracks in their shallow angle, low growth-rate phase,
then turndown of the cracks and the development of trans-
verse defects and rail breaks are prevented.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
880 D. F. CANNON et al.
r
Grinding re-profiles the rail transversely. If the re-profiled
rail is more conformal with the population of wheels that
run over it, contact stresses are lower and crack initiation
is less likely. Re-profiling can also move wheel contact to
areas of the rail less susceptible to RCF.
r
In a rail with existing short cracks, grinding can move con-
tact away from the cracks, thereby stopping their growth.
Although new cracks may be initiated, initiation and initial
growth are slower than propagation of an existing crack.
For practical and economic reasons grinding is not usu-
ally used to remove embryonic cracks that cause shell
formation, or the inclusions that aid initiation. However,
transverse re-profiling of the rail reduces the stresses caus-
ing crack growth. Also a steady removal of metal by grind-
ing continually shifts the area of maximum stress into the
rail, potentially also reducing shell growth.
Although mathematical models now exist to examine
the different stages of crack development, they are not
at a stage, where they can be used to determine robust
and reliable grinding strategies (including metal removal
rates and grinding intervals). Grinding has, however, been
used for more than a decade as a successful treatment
based on a sound understanding of physical principles and
field-testing.
Grinding methods to control RCF
From the 1980s onwards, RCF and the grinding needed
to treat it became increasingly recognized. For example,
Clayton and his colleagues fromBritish Rail Research saw
a need to know, if there was a point at which reducing wear
rates on curves would lead to the emergence of gauge cor-
ner cracking, and argued that regular grinding, to increase
rail-wear rate, could prevent squats.
4143
Further work at
British Rail Research led to the development of a ground
rail profile designed to reduce rolling contact stress.
44
Despite the early British research, until the mid-1990s
grinding in Britain was almost wholly focused on the re-
moval of corrugation. Implementation of grinding to treat
RCF did happen on heavy-haul freight railways. Grinding
was in its infancy in this regard in the early 1980s, when
Table 6 Expectations of rail life on Burlington Northern before and after introduction of preventative grinding
52
1981 1989
Rail Life (MGT) Rail Life (MGT)
Straight track Standard carbon 590680 Standard carbon 7301270
Curves < 1750 m Standard carbon 270320 Standard carbon 7301000
Curves < 870 m Head-hardened 160180 Head-hardened, fully heat treated 450630
Curves < 580 m Head-hardened 110140 Head-hardened, fully heat treated 270450
barely 15% of Canadian Pacific Railways (CPR) grind-
ing budget was devoted to treatment of RCF compared
to 60% on control of corrugation.
45
CPR became one of
the railways in which techniques of regular maintenance
of rails were developed, and CPRs grinding budget now
is devoted overwhelmingly to this end.
At the 1986 International Heavy Haul Association
(IHHA) conference in Vancouver the value of grinding
as a means of curtailing crack growth and modifying the
transverse profile to redistribute loading and reduce con-
tact stresses was noted.
46
Contributions describedwork on
Canadian National Railroad, CPR, and Mount Newman
Mining Railway inAustralia.
4749
By the time of the fourth
IHHA Conference in Brisbane in 1989, the concerns of
heavy-haul railways in general were not with whether to
use rail grinding to control RCF damage, but with how to
grind with greater sophistication.
50
In its earliest use, grinding was used to correct damaged
rail. That is, grinding was used correctively to renovate
rail that was clearly damaged by RCF. In this form of
grinding typically several passes of the grinding train are
needed to rectify the rail. An improvement on this, known
as preventive grinding, was introduced by Kalousek and
his colleagues as a routine maintenance technique.
51
The
objective here is to prevent rail damage rather than to
correct damage. Correctly applied, preventive grinding
can normally be done with a single pass of the grinding
train. Less metal is removed and the logistics of grind-
ing are improved. Grinding intervals are typically in the
range of 1540 MGT. With these techniques of routine
re-profiling and metal removal, the Burlington Northern
railway (now BNSF) was forecasting the increases in rail
life shown in Table 6.
52
Rail lives expected for 1991 have
in fact now been exceeded. Preventative grinding, com-
bined with attention to rail maintenance generally, has
proved to be more successful in extending rail life than
was forecast. On BNSF and elsewhere minimum grind-
ing intervals have been extended, and are now typically
about 2025 MGT even in severe curves.
53,54
Passenger railways have generally been slower to adopt
grinding to treat RCF, despite the fact that RCF (mainly
head-checking and squats) is credited as the cause of about
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 881
1530% of rail breaks on a wide variety of railways. The
severity and widespread extent of the problem in Eu-
rope is evident from a conference in 1992, but, while
research was being undertaken into the problem, rela-
tively little was being learnt from heavy-haul railways re-
garding treatment.
55
The Japanese were less reluctant to
adopt practices on heavy-haul railways, and have under-
taken grinding to treat RCF for more than 10 years. The
problem in Japan would also appear to have been more
severe: in the mid-1990s, fully 70% of the 220 km of an-
nual rail replacement on the Tokaido Shinkansen was for
RCF.
56
In the late 1990s, grinding as a treatment of RCF
of rails began to be adopted on some European rail-
ways.
53,57
It is now fairly widely accepted that the treat-
ments that have been proven elsewhere will also work in
Europe.
Current practices in rail grinding
On heavy-haul railways, it is not uncommon for the length
of track ground per annum to be 50% or more of the
total track length. Some sections may be ground more
than once, and some much less frequently the minimum
interval is in the range 1015 MGT. Most grinding on
such railways is now done to control transverse profile
and provide routine metal removal, reducing RCF and
extending rail life. A recent survey of North American
grinding practices has shown below.
58
r
On the low rails of curves there is near unanimity that
the field side should be relieved to prevent damage and
crushing fromthe false flange of hollowwheels. Acommon
aim is to grind to produce a low rail top radius in the range
180250 mm.
r
On the high rails of curves there is general agreement that
the gauge corner should be ground to give semi-conformal
two-point contact conditions (where the rail quickly wears
to a conformal shape). Heavy gauge corner grinding is no
longer a common practice.
r
In curves, grinding effort tends to be concentrated more on
the low rail (which flattens in service) to achieve a profile
that improves steering.
Grinding is commonly undertaken in a single pass with
a grinding train having about 88 modules, each of more
than 20 kW. Routine, preventive rail grinding has brought
about great increases in productivity: the average grind-
ing speed for Lorams fleet of grinding trains in North
America over a year is about 12 km h
1
(about 8 km h
1
in curved track and 24 km h
1
in straight track), and on
those railways that practice preventive grinding, the aver-
age length of finished ground rail is more than 80%of the
total length of track ground.
In North America, the typical grinding budget for larger
railways is about US$500 per year per kilometre of track,
that is on a system with 20 000 km of track, the grinding
budget is about US$10 million. This figure includes all
costs associated with grinding, not simply the contractors
costs. A general goal of many heavy-haul railways is to in-
crease grinding intervals and reduce the amount of metal
removed through better understanding of rail deteriora-
tion processes.
In Europe, where little track is now ground preventively
for any reason, typically 58% of the total track length is
ground annually, mainly to control corrugation. In British
Railways in the 1990s, less than 1% of the total track
length was ground annually. This had doubled by the end
of the decade and since 2001 has increased greatly in re-
sponse to the acknowledged increase in RCF damage and
the tragic derailment at Hatfield. The volume of grinding
has increased greatly as its value in controlling RCF has
been recognized. This is also being reflected elsewhere in
Europe.
RAI L FAI L URE MODE L L I NG
Much work has been done in the last 3040 years and it
is only possible here to give an indication of the areas of
activity.
Statistical models based on historic data
This simplest type of model primarily depends on the sta-
tistical analysis and presentation of observed behaviour.
This may include defect detection, rail failures, defect
size, inspection performance, derailments etc. Such data
is normally related to some measure of rail loading (e.g.,
months/years of operation) or a global measure of wheel
loading (i.e., gross tonnage). Data are fitted to a statistical
model of some kind and the Weibull distribution is often
found to be the most successful. This type of modelling
is particularly appropriate when specific track conditions,
rail stresses, rail steel fatigue behaviour etc. are unknown
or not sufficiently known; however, the approach can be
developed to study the sensitivity of rail defect behaviour
to such features.
Many models of this kind have been developed and are
often used to relate historical rail-failure data to features
such as rail profile, rail steel type, track curvature. Such
models can be used to identify changes in behaviour and
they may also be used to predict trends. Some very com-
prehensive work of this type has beenundertakeninNorth
America and a good example of it, and how it can be
developed, is presented by Davis et al.
59
This study was
primarily aimed at developing improved rail inspection
programmes but included other technical and economic
aspects.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
882 D. F. CANNON et al.
Mechanistic models
In principle, applying fatigue fundamentals to the initi-
ation and growth of cracks in rails appears straightfor-
ward. In practice it is not, and there are many compli-
cating factors. While static forces are reasonably known,
dynamic forces are highly variable. Track and wheel
discontinuities can lead to high dynamic forces, which
vary with speed and the geometry of the discontinu-
ity. Dynamic forces are also influenced by track geom-
etry, not just whether the track is straight or curved,
but by vertical and lateral irregularities. Other stresses
are present to influence fatigue (as-manufactured residual
stresses and thermal stresses produced by restrained rail
expansion and contraction), and are frequently not well
characterized.
Rail stresses also depend on the amount of wear and the
track structure below the rail (including the sleeper, bal-
last and subgrade condition). In addition, fatigue from
surface or internal defects is dependent on detailed de-
fect geometry, which is often poorly understood. For
example, rail remains in track for many years and is
often subject to corrosion, which can give high stress
concentrations.
With regards to RCF the situation is much more com-
plex. The contact stress field contains high compressive
and shear components, and is very sensitive to the de-
tailed wheel and rail geometries, and to the relative lat-
eral position of the wheel and rail. The relative lateral
wheelrail position in turn depends on the track geom-
etry, the characteristics of the bogie, vehicle speed, the
available adhesion between the wheel and the rail (which
can depend on environmental conditions), and the wheel
and rail geometries. To further complicate modelling, rail
surface cracks grow into a highly strained material with
anisotropic properties that vary with depth, and that is
subject to continuous wear. Crack growth is also encour-
aged by the presence of fluid within the crack that hy-
draulically pressurizes it and reduces crack face sliding
friction.
Prediction of forces from track and vehicle interaction
Mechanistic models require the input of forces that lead
to the stresses mentioned in subsection Rail stresses of
this review. Increasingly, computer-based dynamic models
are used to predict vehicle/track interaction; these include
ADAMS/Rail, MEDYNA, Vampire

, GenSys, and NU-


CARS
TM
packages that have been developed by research
institutions and railway administrations around the world.
When rail failures associated with the wheelrail contact
surface are being considered it is necessary to calculate
local conditions and this is often done with sub-packages
which either assume Hertzian or non-Hertzian contact
conditions.
Fatigue crack initiation
Away fromthe rolling surface of the rail, fatigue crack ini-
tiation can be predicted using classic methods involving
the concepts of strain/stress life relationships, monotonic
and cyclic stress/strain relationships, stress or strain con-
centration factors and Miners damage summation rule for
variable amplitude loading.
At and near the rail running surface modelling is more
complex. Recent studies of this modelling area include
those at the University of Sheffield and Chalmers Univer-
sity, Gothenburg.
60,61
This area of work has the greatest
potential to support the development of strategies to com-
bat RCF problems. It has application to track and vehicle
design, material development and evaluation, optimized
wear and/or grinding rates, and rail replacement plan-
ning. Given the modelling problems, validation is critical.
Some work on this has been undertaken in the recent UIC
Rail Defect Management Project.
1
Fatigue crack propagation
Crack propagation is universally dealt with by the applica-
tion of fracture mechanics. Some rail failure models that
incorporate fatigue crack propagationare discussed below.
The Association of American Railroads PHOENIX
model contains a rudimentary crack growth model (a
penny shaped crack) for the small transverse growth of
a shell to form a transverse defect.
62
Direct and shear
stresses are accounted for by employing an effective
range of the crack tip stress intensity factor K.
Some verification of the PHOENIX model has been
possible using data from the FAST facility at the FRAs
Transportation Technology Centre in Pueblo, Colorado,
and the model has been used to examine, for example,
the effect of wheel load, grinding and rail profile on shell
fatigue life.
A model that considers the mechanism by which a hor-
izontal shell turns down into a transverse detail fracture
(Fig. 13) has been developed by Farris et al.
63
The turn-
ing phase of relatively benign horizontal head cracks is
particularly significant with respect to rail life, inspection
criteria andmethods, andactioncriteria. The Farris model
is based on a two-dimensional, horizontal and frictionless
subsurface crack. Contact stresses and residual stresses are
imposed and create crack tip stress intensity factors at each
end of the crack. Direct and shear stresses lead to mode I
and II factors. However, only the reversed mode II values
are considered to be relevant to propagation and turning.
Propagation is related to the range of K
II
and turning is
related to the sum of the maximum and minimum mode
II stress intensities. This model has been used, for exam-
ple, to predict the effect of residual stresses, shell depth
and rail running surface coefficient of friction on turning
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 883
Fig. 13 Transverse crack originating from a horizontal shell.
behaviour. Orringer et al. have developed similar models
for the detail fracture.
64
The above models are typical of those that deal with sim-
ple loading histories; however, in reality the conditions
in which cracks develop close to the rail running surface
(cracks such as squats and head checks) are rather dif-
ferent. Extreme plastic deformations occur during stress
histories that are variable in magnitude and complex in
their sequential characteristics. Fatigue crack growth in
rail steels under mixed, sequential conditions has been
studied, by, amongst others, Brown and Bold.
65
Two- and three-dimensional finite element modelling
of RCF-type cracks has been undertaken.
66
Bringing
together laboratory mixed-mode fatigue crack growth
rate studies and the finite element modelling of surface-
initiated RCF cracks is leading to the development of
models, which can predict crack behaviour in a mix of
wet and dry conditions.
67
Such modelling is beginning
to indicate the significance of fluid entrapment and op-
erational parameters on the propagation of RCF cracks.
However, much more experimental mixed mode fatigue
data are still required.
Finite element modelling of RCF by Schnitzer indicates
that when the crack is closed by bending forces the wheel
rail contact stresses create a dominating cyclic mode II
loading.
68
It is claimed that the cracks direction, particu-
larly its tendency to turn from a shallow to a steeper angle
(with respect to the running surface) at a depth of about
5 mm, is due to a change from orthotropic fracture me-
chanics properties in the near surface layer of cold worked
steel to isotropic fracture mechanics. The change in crack
direction leads to either spalling of the surface or trans-
verse cracking and possibly rail fracture.
Modelling of fatigue crack growth provides a major input
to the development of inspection strategies and the defini-
tion of actionable defect sizes. Crack growth rates in many
families of steels have been found to be similar; however,
other dissimilar properties can affect residual stress fields
and these can have a substantial affect on growth rates.
Fracture
As with fatigue crack propagation, almost all modelling
of rail fracture has been based upon fracture mechanics.
In most cases it has been reasonable to assume that sim-
ple mode I crack opening conditions apply (or dominate)
at the point of fracture. However, temperature and, to a
lesser extent, loading rate sensitivities have to be taken
into consideration. Critical crack lengths need to be de-
fined, but it is often the case that fatigue crack growth
rates are so high in the later stages of crack development
that precise sizes are not too important.
Probabilistic simulation of rail failure
A simple approach to this aspect of modelling is reliability
analysis. Eisenmann has used a classic reliability analy-
sis to assess the probability of failure of plain corroded
rail and alumino-thermic welds.
69
Such analysis requires
at least two probability density functions; one describing
the loads applied and the other describing the strength
or resistance of the loaded body. The overlap of the two
functions is used to describe a third unreliability distri-
bution, which can be used to define probability of failure
and a safety index. Eisenmanns example considers fail-
ure from the rail foot and the loading distribution is es-
sentially that of wheelrail force translated into bending
stress. The strength distribution is provided by fatigue
test data distributions for corroded UIC 60 section rail
and alumino-thermic welds. Comparing the performance
of plain corroded rail to the aluminothermic weld, Eisen-
mann concludes that the latter does not particularly create
a weak link in the track system.
Reliability analysis can be taken further by extending the
content of the loading and strength distributions. Some-
times additional variables can be readily mathematically
combined, adding to the loading and strength distribu-
tions. However, in many cases this is not readily possible
as the distributions to be modelled may be of an arbitrary
form. In such cases the Monte Carlo simulation method is
often used to combine variables. Large numbers of analy-
ses are performed, for example, using a rail-failure model,
and in each case discrete values of the variables in the
model are randomly chosen from the (arbitrary) distribu-
tion hence the name Monte Carlo. The method is not
new but it lacked popularity until the advent of power-
ful and fast computing, which enables many calculations
to be done so that predicted distributions stabilize in a
reasonable length of computing time.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
884 D. F. CANNON et al.
Monte Carlo simulation can be used to examine crack
growth behaviour, critical crack sizes, and the determina-
tion of allowable defect sizes in rails.
7072
It conveniently
enables inspection parameters, risks and costs to be intro-
duced into rail-failure models. An example of the latter is
work to assess the so-called delayed-action concept for
rail inspection.
73
In this, defined inspection/action proce-
dures allowdetected sub-critical defects to remain in track
for some period of time until a further inspection is carried
out or some other action is taken. The work of Ref. [73]
included three basic models: a fracture-mechanics model
to predict growth of an internal rail-head crack; a model
to simulate rail flawdetection and rail removal; and a risk
benefit analysis. In this example, the Monte Carlo simula-
tion indicated potential economic benefit from using the
delayed-action concept.
Suchdelayed actionis commonly applied inEurope. The
applicationof probabilistic fracture mechanics was studied
by the European Railway Research Institute.
70
Using this
work as a basis, limit sizes for defects requiringrail removal
immediately and within 4 weeks were determined for the
former East German Railways.
71
CONCL UDI NG COMME NT S
This review has tried to cover many, but not all, topics
relating to railway rail failures and inevitably much work
has been omitted. Despite this, it is hoped that the review
presents an impression of where we are today and how we
got there.
The emergence of surface-initiated rail RCF as a major
cause of premature rail removal is of great concern as it
indicates that operating conditions are taking the rail to
and beyond its natural endurance limit. This review indi-
cates that current research and modelling activity are very
much focussed on this issue, but the problem is complex
and much still must be done. It is likely that a major step
in rail/wheel technology will be required to solve the RCF
problem. Until that time, RDM and maintenance strate-
gies can be developed that should contain the current rail
failure situation.
Despite major improvements in rail making and inspec-
tion, rail breaks still occur: for example, in the UK the
annual number of broken rails remained almost constant
at about 770 per year between 1969 and 2000. Rail fail-
ures began with the birth of the industry in the early
years of the industrial and technological revolution. In the
past, many railways have, consciously or unconsciously,
believed that some level of broken rails is unavoidable.
The costs of reducing rail failures have to be balanced
against reduced costs from death and injury, penalty pay-
ments, and train disruption, increased customer satisfac-
tion, and better planning of track maintenance and re-
newal. It should also be recognized, that rail failure and
vehicle derailment can lead to a loss of public confidence
that has devastating and long-lasting effects on the rail
industry. Rail failure should not be seen as inevitable. It
is essential that research institutions and the international
rail supply and operating businesses collaborate and vigor-
ously pursue a strategy with the long-termaimof reducing
rail failures to zero.
RE F E RE NCE S
1 Lundgren, J. R., Cannon, D. F. and Zuber, P. (2001) An
international cooperative research approach to rail defect risk
management. In: Seventh International Heavy Haul Association
Conference, Brisbane, Australia.
2 Temming, R. L. (2001) Das groe Eisenbahnbuch. Neuer Kaiser,
Verlag GmbH, Klagenfurt.
3 Herring, P. (2000) Ultimate train. In: Deutschsprachige Ausgabe:
Die Geschichte der Eisenbahn. Dorling Kindersley Limited,
London.
4 Rossberg, R. R. (1935) Geschichte der Eisenbahn, Sigloch
Edition, K unzelsau, ohne Jahr. ISBN: 3-89393-174-0; p. 26,
p. 35.
5 Hundert Jahre deutsche Eisenbahnen. Herausgegeben von der
Hauptverwaltung der Deutschen Reichsbahn.
6 Sperry Rail Service (1989) Rail Defect Manual. Danbury,
Connecticut, USA.
7 European Standard, Final Draft prEN 13674-1 (November
2002) Railway Applications Track Rail, Part 1: Vignole
railway rails 46 kg/m and above.
8 Edel, K. O. (1979) Temeratureabh angigkeit des Auftretens von
Schienenbr uchen. Signal Schiene 23, 266272.
9 Krabiell, W. and Dahl, W. (1982) Zum Einflu von Temperatur
und Beanspruchungsgeschwindigkeit auf die Riz ahighkeit von
Baust ahlen mit unterschielicher Festigkeit. Archiv. F ur
Eisenh uttenwesen 53, 225230.
10 Cannon, D. F. and Allen, R. J. (1974) The application of fracture
mechanics to railway failures. I. Mech. E. Rail. Engng. J. 3, 6
17.
11 Edel, K. O. (1987) Untersuchung des Bruchverhaltens von
Eisenbahnschienen und vollr adern. Technische Hochschule
Otto von Guericke Magdeburg, Dissertation B.
12 T oth, L. (2000) Fracture properties of rail steels and their
practical use. In: Internationales Symposium Schienenfehler, Bran
den burg an der Havel, Tagungsbericht Seite 12-1 bis 12-12.
13 Davis, D. D., Sawley, K. and Guillen, D. G. (2001) Bainitic
steel frogs and rails show promise. Int. Railw. J., XLI, 1920.
14 Orringer, O., Tang, Y. H., Gordon, J. E., et al. (1988) Crack
propagation life of detail fractures in rails. US Department of
Transportation, FRA, DOT/FRA/ORD-88/13.
15 Ringsberg, J. W. and Josefson, B. L. (2000) Finite element
analyses of rolling contact fatigue crack initiation in railheads. I.
Mech. E. J. Railw. Rapid Transit. 215, 243259.
16 ERRI (1995) Measurement of Residual Stresses using Ultrasonics.
European Railway Research Institute Report D173/RP12.
17 ERRI (1993) Measurement of Residual Stresses under the Running
Surface of the Rail by Neutron Diffraction. European Railway
Research Institute Report D173/RP4.
18 Bower, A. F. and Johnson, K. L. (1991) Plastic flow and
shakedown of the rail surface in repeated wheel-rail contact.
Wear, 144, 118.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
RAI L DEFECTS: AN OVERVI EW 885
19 Bogdanski, S., Stupnicki, J., Brown, M. and Cannon, D. F.
(1997) A two-dimensional analysis of mixed-mode rolling
contact fatigue crack growth rates in rails. In: Fifth International
Conference on Biaxil/Multaxial Fatigue and Fracture, Cracow.
20 ERRI (2000) Rail Defect ManagementState of the Art. European
Railway Research Institute Report D229/RP2 (2nd edn).
21 Rosche, H. (1906) Ausf uhrung und Unterhaltung des
Oberbaues. In: Der Eisenbahnbau, Zweiter Band: Berechnung,
Konstruktion, Ausf uhrung und Unterhaltung des Oberbaues.
(Edited by F. Loewe, H. Zimmermann (Herausgeber)),Verlag
von Wilhelm Engelmann, Leipzig, 2.
22 ORE (1968) Untersuchung der Schienenfehler im Gleis Bericht
Nr.: Vereinheitlichte Statistiken uber das Auswechseln von
Schienen im Jahre 1965 Schienenbr uche, die zu Entgleisungen
gef uhrt haben. In: Forschungs- und Versuchsamt des Internationalen
Eisenbahnverbandes. Utrecht, ORE-Bericht D88/RP 5.
23 NTSB (1974) Broken Rails: A Major Cause of Train Accidents.
U.S. National Transportation Safety Board, Special Study
NTSB-RSS-74-1, 2.
24 Resor, R. (1997) Broken Rail Detection with Current Technology.
Workshop on Rail Defect and Broken Rail Detection. TTCI, Pueblo.
25 Dearden, J. (1957) Rail failures on british railwaysIncidence,
and preventive measures. Railw. Gazette 113, 509512.
26 Dearden, J. (1965) Rail failures on british railwaysreporting,
analysis and prevention. Railw. Gazette 121, 148150.
27 Herwig, V. (1939) Steel rail failures. Railw. Gazette 95, 5256.
28 Brite/Euram (19971999) Integrated Study of Rolling Contact
Fatigue (ICON)European Commission DG X11 Brite/Euram III
Project. Contract BRPR-CT96-0245 Project Programme,
Brussels.
29 UIC (2002) Handbook of Rail Defects. UIC Code 712R (4th
edn), Paris.
30 OSShD (1961) Empfehlungen uber die Festlegung der
einheitlichen Klassifikation der Fehler und Besch adigungen an
Schienen, Organisation f ur die Zusammenarbeit der
Eisenbahnen (OSShD) OSShD-Merkblatt R 733, 1.
31 OSShD (1985) Empfehlungen uber die Klassifikation der
Schienenfehler. Organisation f ur die Zusammenarbeit der
Eisenbahnen (OSShD). Merkblatt R 733, 2.
32 UIC/ERRI (2001) International Cross Reference of Rail
Defects. UIC/ERRI Document D229/RP3.
33 Hogenkamp, F. (1981) Zerst orungsfreie Pr ufung mit
Ultraschall an Schienen und Weichen im Gleisein

Uberblick.
Eisenbahningenieur 32, 375382.
34 Hug, H. and Meiner, K. (2000) Detektion und Behandlung
von Schienenfehlern in Gleisen und Weichen der DB Netz AG.
In: Internationales Symposium Schienenfehler. Fachhochschule
Brandenburg, Tagungsbericht.
35 Kecskes, S. (1972) Ermittlung der Entwicklung von
nierenf ormigen Erm udungsrissen im Schienenkopf in den
Linien der Ungarischen Staatseisenbahnen. Schienen der Welt 3,
545559.
36 Poro sin, V. L. (1980) Vlijanije ekspluatucionnych faktorov na
skorost razvitija popere cnych tre s cinv golovke relsov. Vestnik
VNII

ZT 39, 4547.
37 Mair, R. I. and Groenhout, R. (1981) Das Anwachsen von
Erm udungs-Querfehlern in Eisenbahnschienenk opfen. Schienen
der Welt, 12, 179196.
38 Krull, R., Hintze, H. and Thomas, H.-M. (2000) Moderne
Methoden der zerst orungsfreien Werkstoffpr ufung im
Oberbau. In: Internationales Symposium Schienenfehler,
Brandenburg an der Havel, 3-13-16.
39 AREMA (2001) American Railway Engineering and
Maintenance-of-Way Association Manual for Railway Engineering.
Vol. 1 (Track), Section 4.6.
40 Grassie, S. L. and Kalousek, J. (1997) Rolling contact fatigue of
rails: characteristics, causes and treatments. In: Sixth
International Heavy Haul Association Conference, Cape Town,
South Africa, 381404.
41 Clayton, P., Allery, M. B. P. and Bolton, P. J. (1983) Surface
damage phenomena in rails. In: Proceedings Conference on Contact
Mechanics and Wear of Rail/Wheel Systems. (Edited by J.
Kalousek, R. V. Dukkipati and G. M. L. Gladwell). University
of Waterloo Press, 419443.
42 Clayton, P. and Allery, M. B. P. (1982) Metallurgical aspects of
surface damage problems in rails. Can. Metall. Q. 21, 3146.
43 Clayton, P., Allery, M. B. P. and Bolton, P. J. (1982) Surface
damage of rails. Rail Technology, (Edited by C.O. Frederick and
D. J. Round) Nottingham, UK,179192.
44 Smallwood, R., Sinclair, J. C. and Sawley, K. J. (1991) An
optimisation technique to minimise rail contact stresses. Wear
144, 373384.
45 Roney, M. D. (1983) Economic aspects of the wear of rail on
canadian railways. In: Proceedings of Conference on Contact
Mechanics and Wear of Rail/Wheel Systems. (Edited by J.
Kalousek, R. V. Dukkipati and G. M. L. Gladwell) University of
Waterloo Press. 271291.
46 Proceedings of Third International Heavy Haul Association
Conference, Vancouver, 1986.
47 Worth, A. W., Hornaday, J. R. and Richards, P. R. (1986)
Prolonging rail life through rail grinding. In: Third International
Heavy Haul Association Conference. Vancouver, 106116.
48 Lamson, S. T. and Roney, M. D. (1986) Development of rail
profile grinding on cp rail. In: Third International Heavy Haul
Association Conference, Vancouver, 98105.
49 Epp, C. J. (1986) Wheel and rail profiling to control system
performance. In: Third International Heavy Haul Association
Conference, Vancouver, 7582.
50 Proceedings of Fourth International Heavy Haul Association
Conference, Brisbane, 1989.
51 Kalousek, J., Sroba, P. and Hegelund, C. (1989) Analysis of rail
grinding tests and implications for corrective and preventative
grinding. In: Fourth International Heavy Haul Association
Conference, Brisbane, 193204.
52 Glavin, W. E., Aspebakken, J. I. and Besch, G. O. (1989) Heavy
haul: The Burlington Northern perspective. In: Fourth
International Heavy Haul Association Conference. Brisbane,
276285.
53 Grassie, S. L., Nilsson, P., Bjurstrom, K., et al. (2000)
Alleviation of rolling contact fatigue on swedens malmbanan.
In: Cm2000: Fifth International Conference on Contact Mechanics
and Wear of Rail/Wheel Systems. Tokyo.
54 Stanford, J., Magel, E. and Sroba, P. (2001) Transitioning from
corrective to preventive grinding on the BNSF railroad. In:
Seventh International Heavy Haul Association Conference, Brisbane,
493501.
55 Kalker, J. J., Cannon, D. F. and Orringer, O. (Eds) (1993) Rail
Quality and Maintenance for Modern Railway Operation, Delft
(1992) Kluwer Academic Publishers, Dordrecht.
56 Kondo, K., Yoroizaka, K. and Sato, Y. (1996) Cause, increase,
diagnosis, countermeasures and elimination of shinkansen
shelling. Wear 191, 199203.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886
886 D. F. CANNON et al.
57 Grohmann, H.-D. and Schoech, W. (2000) Contact geometry
and surface fatigue minimising the risk of head-check
formation. In: Cm2000: Fifth International Conference on Contact
Mechanics and Wear of Rail/Wheel Systems. Tokyo.
58 Sawley, K. (1999) North American Rail Grinding: Practices and
Effectiveness. Association of American Railroads Report
R-928.
59 Davis, D. D., Joerms, M. J., Orringer, O. and Steele, R. K.
(1987) The Economic Consequences of Rail Integrity. Association of
American Railroads Report R-656.
60 Beynon, J. H., Brown, M. W., Joyce, R. D., et al. (2000)
Integrated study of fatigue and wear of pearlitic rail steels under
rolling-sliding contact conditions. In: Cm2000: Fifth
International Conference on Contact Mechanics and Wear of
Rail/Wheel Systems. Tokyo.
61 Ringsberg, J. W. (2001) Life prediction of rolling contact
fatigue crack initiation. Int. J. Fatigue 23, 575586.
62 Steele, R. K. and Muhlemburg, J. A. (1992) Validation of Rail
Fatigue Life, Crack Growth and Deformation Behaviour Predictions.
Association of American Railroads Report WP-153.
63 Farris, T.N., Keer, L. M. and Steele, R. K. (1991) Life
prediction for unstable shell growth in rails. In: Mechanics and
Fatigue in Wheel/Rail Contact, Elsevier.
64 Orringer, O., Tang, Y. H., Gordon, J. E., Jeong, D. Y., Morris,
J. M. and Perlman, A. B.. (1988) Crack Propagation Life of Detail
Fractures in Rail. Final Report: DOT/FRA/ORD-88/13.
65 Brown, M., Hemsworth, S., Wong, S. L. and Allen, R. J. (1996)
Rolling contact fatigue crack growth in rail steel. In: Second
Mini-Conference on Contact Mechanics and Wear of Rail/Wheel
Systems, Budapest.
66 ERRI (1997) Stress Analysis of Rail Rolling Contact Fatigue Cracks.
European Railway Research Institute Final Report D173/RP19,
Utrecht.
67 Bogdanski, S., Stupnicki, J., Brown, M. W. and Cannon, D. F.
(1997) A two dimensional analysis of mixed-mode rolling
contact fatigue crack growth in rails. In: Fifth International
Conference on Biaxial/Multiaxial Fatigue and Fracture,
Cracow.
68 Schnitzer, T. (2000) Bruchmechanische analyse der
Riausbreitung durch Rollkontakterm udung. In: Internationales
Symposium Schienenfehler. Fachhochschule Brandenburg,
18-118-20.
69 Eisenmann, J. (1996) Ausfallwahrscheinlichkeit von
Schienenquerbr uchen bei l uckenlos verschweiten Gleisen.
ETR, 45.
70 Edel, K.-O. (1987) Die Festlegung zul assiger Rigr oen f ur
Schienen und Schienenschweiungen auf der Grundlage der
probabilistischen Bruchmechanik. In: Forschungs- und
Versuchsamt des Internationalen Eisenbahnverbandes. Utrecht,
Technisches Dokument DT 183 (E 162).
71 Edel, K.-O. (1990) Grenzmae f ur Risse in Eisenbahnschienen
der DR. Signal Schiene 34, 212214.
72 Edel, K.-O. (1991) Sicherheitsbewertung des Riverhaltens. In:
Deutscher Verband f ur Materialfor schung und -pr ufung e.V.,
Vortr age der Tagung Werkstoffpr ufung 1991, Bad Nauheim,
8392.
73 Jeong, D.Y., Orringer, O., Tang, Y. H. and Perlman, A. B.
(1997) Evaluations of rail inspection programs. In: Rail Defect
and Broken Rail Detection: Workshop on Rail Defect and Broken Rail
Detection, TTCI, Pueblo, Colorado, USA.
c
2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 865886

You might also like