You are on page 1of 32

12

Rail welds
M. J. M. M. Steenbergen, Delft University of Technology,
The Netherlands; R.W. van Bezooijen, Id2 Consultancy,
The Netherlands

Abstract: Welding is the standard for joining rails in modern railways.


This chapter discusses several aspects of rail welding. The basic welding
technologies are presented, including metallurgical aspects and post-welding
treatments. Different damage mechanisms initiated at or by rail welds are
addressed. The dynamic wheel–rail interaction at rail welding irregularities
is investigated, leading to quantitative relationships between key parameters
describing this interaction. A method for force-based weld assessment is
elaborated as an alternative for the conventional method based on tolerances.
The chapter concludes with energy considerations in relation to welding
irregularities and track deterioration.

Key words: rail welds, weld geometry, dynamic wheel–rail interaction, weld
assessment, track deterioration.

12.1 Introduction
Continuously welded rail (CWR) was introduced on the railways in the
1930s, and nowadays is the standard in modern railway tracks. In many
cases, the traditional bolted or fish-plated rail connections are present in the
track only in the form of insulated rail joints for detection and signalling
purposes. The deflection of a bolted rail joint under static train axle loading
leads to a high dynamic impact component, which is a source of rapid
track deterioration and high noise levels (Kabo et al., 2006; Steenbergen,
2006; Cai et al., 2007). Therefore, the welded continuous connection was
a significant improvement.
In the present chapter, the main different rail welding technologies will be
discussed briefly in Section 12.2. After that, damage mechanisms that may
be initiated at or by rail welds are addressed in Section 12.3. An important
issue is the dynamic wheel–rail interaction at rail welds; this is the subject
of Section 12.4. In Section 12.5 the Dutch rail welding regulations will be
explained as a method for force-based weld assessment, as an alternative
to the conventional method purely based on geometry. Section 12.6 will
conclude with energy considerations in relation to welding irregularities and
track deterioration.

377
378 Wheel–rail interface handbook

12.2 Rail welding processes


The most common types of rail welding are electrical flash-butt welding and
aluminothermic or thermite welding (Fig. 12.1). Enclosed arc welding is applied
only in specific situations such as making closures. The application of the

(a)

(b)

12.1 Flash butt (a) and aluminothermic (b) rail welding on the Dutch
HSL-South.
Rail welds 379

first type of welding is almost fully automated, including the positioning of


the rail ends to be joined, and it is therefore often applied in the construction
of new tracks or large renewal projects. It can be applied in stationary plants
as well as mobile machines. The second welding process requires handicraft
and is therefore often applied in repair and small maintenance works, or
installation of insulated rail joints, switches and crossings. Both rail welding
procedures are explained below.
The mobile flash-butt rail welding process (Wöhnart and Wenty, 2002)
begins with the clamping of the rail ends by jaws in the welding unit. They
act as water-cooled electrodes that ensure the transfer of the welding current
to the rails. The rail ends are aligned and brought together by a hydraulic
movement of the two aggregate halves of the welding unit, thus producing
arcing. Now follows the welding process in three stages: pre-heating, a stage
with constant pressing together of both rail ends and a progressive stage of
accelerated burn-off. During this process, current intensity and flash-burn
speed depend on the stage and the rail grade. The welding process ends with
a final compressing stroke, with the aim to avoid welding nuggets, oxide
inclusion and impurities from flashing. After welding, shearing knives,
following the rail cross-sectional geometry, encompass the rail and trim the
welding seam while still red hot.
Special alloy and head-hardened rails require special processing during
the cooling stage. Special alloy rails should be prevented from cooling
down too quickly and follow a post-heating program in order to prevent the
development of brittle martensite parts. On the other hand, head-hardened
rails need to be cooled down quickly in order to keep the hardness level of
the parent material intact. This can be achieved by ‘air quenching’, yielding
hardening of the railhead.
The aluminothermic welding method is a fusion welding process and
makes use of a mixture of aluminium powder and iron oxide which is
converted into alumina and steel at a high temperature (about 2500 ºC).
The process is as follows: both rail ends are aligned; a mould is installed
around the gap in the joint (about 20–25 mm) with a crucible on top; the
rails are pre-heated to about 1000 ºC using propane burners; the mixture in
the crucible is ignited, yielding an exothermic chemical reaction and filling
the gap within the mould; then both crucible and mould are removed, and
the weld collar is stripped and ground.
The alignment (a small overheight of about 1.2–1.4 mm is given) of the
rail ends for both welding methods is done in order to account for the unequal
contraction of the head and the foot of the rail profile after welding during
cooling down; in this way ‘dipped’ welds are avoided. Cooling down after
welding requires special attention in order to maintain the pearlitic crystalline
rail structure. Special so-called TTT (time–temperature transformation)
diagrams provide relationships between cooling rates and steel phases for
380 Wheel–rail interface handbook

different rail steel grades and alloys, such that austenite formed during rail
heating is re-transformed into pearlite and the formation of brittle martensite
is avoided. The manual thermite welding process in particular requires
craftsmanship, in order to control and optimize the steel phases and their
specific properties with pre-heating, controlled cooling down and post-weld
treatment under all circumstances and conditions.
Generally, the fatigue resistance of flash-butt welds is superior to that
of aluminothermic welds; estimations for heavy-haul traffic (Union Pacific
Railroad) are 365 MGT for thermite welds against 877 MGT for flash-
butt welds (Wimmer et al., 2002). In the case of severe welding geometry
irregularities, the fatigue life of welds may be significantly affected by the
effect of the dynamic wheel load and the occurrence of loose sleepers (Ishida
et al., 1999; Steenbergen, 2008b).
In order to obtain a good geometry, the final grinding of the rail surface
is in most cases done manually for both types of welding. In some cases,
however, the welds are treated using a grinding train, on newly built tracks,
yielding a high surface quality (Fig. 12.2).

12.3 Rail welds and damage formation


The rail weld is, in comparison to plain rail, particularly susceptible to
damage formation. This is mainly due to three reasons. First, the presence
of the weld yields a geometrical irregularity along the wheel–rail interface,
leading to dynamic axle load variations and dynamic contact stress
amplifications. This concerns both normal stresses and tangential stresses
and their distribution in the rolling contact. A second reason is the presence
of material inhomogeneities along the wheel or rail surface. This includes
a non-constant hardness distribution along the rail and microstructural
disturbances, non-metallic inclusions or steel phase differences along the
surface or subsurface. A third reason is the presence of internal material
inhomogeneities or clusters of them in the weld; the microcleanliness of
the weld metal in cast condition is generally inferior to that of the parent
material. In general, none of the previous mechanisms occurs independently
when leading to damage: they interact, significantly increasing the propagation
speed of existing damage.
It is obvious from the above that the aluminothermic weld in particular is
susceptible to these phenomena. The geometrical irregularity at the rail weld
which is present in the majority of the cases can be due to many reasons.
These include the effects of unstraightened rail ends, inaccurate rail end
positioning, the manual grinding of the rail surface and the influence of
inhomogeneous material shrinkage after cooling down in combination with
early grinding. The length of the irregularity is generally less than 1 m.
Therefore, in general the rail welds yield a significant contribution to the
Rail welds 381

(a)

(b)

12.2 Manual rail weld grinding (a) and grinding train (Speno) (b) on
the Dutch HSL-South.

short-wave contribution of the track irregularity spectrum, which determines


the rate of the rail damage progress and track deterioration: dynamic forces
may damage the rail or railhead, the concrete sleepers and rail fastenings
382 Wheel–rail interface handbook

and influence, for example, wheel seat fretting fatigue; furthermore, they
are responsible for non-uniform ballast settlement or local decompaction.
It is important to note that, especially on newly built tracks, the short-wave
contribution is almost entirely determined by the rail weld geometry.
Normally, new welds are being tested ultrasonically for internal damage,
cracks, voids or inclusions before acceptance. In general this is combined
with magnetic particle or eddy current inspection for the lop layer (5–7
mm). When a not-detected serious internal material defect, for example a
lack of fusion or porosity, is present in a weld, it often breaks at an early
stage (Shitara et al., 2003; Terashita and Tatsumi, 2003), especially in cold
winters with high tensile forces in the rail. Therefore, internal material or
fusion defects are in general not responsible for long-term deterioration at
the wheel–rail interface.
The effect of locally varying metallurgical properties along the rail surface
however is particularly relevant for rail welds. As has been mentioned, often
short indentations along the rail surface occur at the centre of rail welds, at
the transitions from the weld material to the parent material, the so-called
heat-affected zone (HAZ) (Fig. 12.3). The phenomenon is due to shrinkage
after grinding and further cooling down, especially when the steel temperature
was not low enough at the moment of final grinding. It can be observed
particularly for aluminothermic welds in rail repair works, executed under
time pressure and non-optimal conditions.
These geometrical indentations typically coincide with the local hardness
minima along the rail surface in the HAZ at both sides of the weld metal
(Mutton, 2000). This is shown qualitatively in the diagram in Fig. 12.4.
Vickers hardness levels (at 5 mm below the surface) typically fluctuate
between 250 and 400 HV. For comparison, the hardness of the parent rail
generally starts around 260 HV for new common rail grades (increasing to
300 HV after grinding and initial cold work-hardening due to train operation)
up to 400 HV for new head-hardened rails.
The unevenness in the rail surface, resulting in a local fluctuation of contact
stresses in the rolling contact, may induce differences in shakedown, work-
hardening and plastic deformation of the rail surface along the rail (Böhmer
and Klimpel, 2002), leading to non-homogeneous wear, surface cracking and
a progressively deteriorating geometry. This process is also called weld batter
(Mutton and Alvarez, 2004). Examples of this phenomenon are shown in Fig.
12.5. In in-situ welding it is almost impossible to avoid these indentations,
except for new track construction works, where the grinding can be done
separately from the welding.
Due to the effects that have been discussed previously, rail welds are often
considered, and also observed in practice, as corrugation or squat initiators
(Hiensch et al., 2002; Li et al., 2006, 2007). In Figs 12.6 and 12.7 examples
are shown of squats on aluminothermic and flash-butt welds, respectively.
Rail welds 383

The residual stress field which is induced over the cross-section of the
rail as a result of welding and heat transfer processes may be a further cause
of internal damage propagation or failure (Chen et al., 2006a,b). Webster
et al. (1997) showed that in aluminothermic welds the longitudinal residual
stresses are compressive at the top and the bottom of the rail, which is

(a)

(b)

12.3 (a) Weld and parent material; (b) the HAZ after thermite welding
and grinding.
384 Wheel–rail interface handbook

Parent material HAZ Weld metal HAZ Parent material

12.4 Qualitative behaviour of the steel hardness distribution along


the rail surface.

beneficial for the fatigue lifetime. In the web region, however, residual stresses
are strongly tensile, which makes that region very susceptible to damage
initiation and growth. Tawfik et al. (2006) performed a similar analysis for
flash-butt welds, and showed again a strong residual tensile stress field in
the web of the rail, leading to specific failure modes originating from the
rail web. Since the rail web mostly accounts for the shear stresses in the
cross-section, these are failure modes induced by shearing under high axle
loads, such as the horizontal split web mode. Tawfik et al. (2008) showed
that the tensile residual stress field may be alleviated by localised post-weld
heat treatment at the base of the foot directly after welding. Without heat
treatment, stress values in longitudinal direction may be as high as 300
MPa. Skyttebol et al. (Skyttebol, 2004; Skyttebol et al., 2005) showed that
the residual stress field in combination with stresses resulting from ambient
temperature loading and train axle loading may lead to very rapid crack or
damage growth and a significantly reduced fatigue life. Mutton and Alvarez
(2004) observed that the fatigue process leading to the horizontal split web
failure mode on heavy-haul lines (especially on curves and tangent track)
is in many cases low-cycle fatigue with few high axle overloads and rapid
fracture growing from weld collar slag inclusions. Ilic et al. (1999) showed
experimentally the effect of post-weld heat treatment also on the crystalline
structure of thermite welds (grain size reduction and homogenizing) leading
to higher ductility.

12.4 Rail welding irregularities and dynamic


effects in the wheel–rail interface
As has been discussed previously, the rail surface is in general not perfectly
smooth in the presence of rail welds, giving rise to dynamic wheel–rail
interaction in the contact patch. In Fig. 12.8 (Steenbergen and Esveld, 2006b),
finite element (FE) simulation results are shown for a train vehicle, passing
an artificial irregularity in the track at a speed of 140 km/h. The irregularity is
typical for a rail welding irregularity: it comprises both a smooth irregularity
with a longer length-scale (a harmonic wave with a length of 1 m and a top
Rail welds 385

(a)

(b)

12.5 Deteriorating geometry resulting from non-uniform shakedown


and work hardening along a rail weld in the HAZ; (a) flash-butt weld;
(b) thermite weld.
386 Wheel–rail interface handbook

(a)

(b)

12.6 Squat initiation on thermite rail welds or in the HAZ (a);


accompanied by crack initiation (b).

value of 1 mm), typically due to unstraightened rail ends or the upset of


rail ends before joining, and a second non-smooth irregularity with a short
length-scale (a triangular peak with a basis of 100 mm and a top of 0.2 mm),
typically due to welding or shrinkage.
Rail welds 387

(a)

(b)
12.7 Squat initiations on flash-butt rail welds or in the HAZ.

The results in Fig. 12.8 have been obtained from FE simulation with the
DARTS-NL package. Parameter values were adopted as follows: a wheel
mass (half unsprung mass) of 970 kg; a sleeper mass of 300 kg; the 54E1
rail profile is used; the primary suspension stiffness equals 1.8.106 N/m (per
wheel); the railpad stiffness 1.2.109 N/m and the ballast stiffness 30.106 N/m
per sleeper. The latter value is rather small, to account for the generally bad
compaction of the ballast underneath the sleepers close to the irregularity;
388 Wheel–rail interface handbook

due to ‘ballast pumping’ loose sleepers often occur (Ishida et al., 1999). The
package uses a non-linear Hertz contact model.
In Fig. 12.8 the following quantities are shown as a function of time:
• the geometry of the irregularity z;
• the dynamic component of the vertical wheel displacement of the first
wheel of a passing bogie, defined positive in upward direction and
calculated in a convective reference frame moving along with the
wheel;

1.2
0.8 Artificial weld irregularity 1 m
z [mm]

0.4 V = 140 km/h


0
0 0.04 0.08 0.12
x /V [s]

1
uwheel/utrack,dyn [mm]

0.6

0.2
–0.2
0.04 0.08 0.12
–0.6 t [s]
Wheel displacement
–1 Track displacement
120
P1
Fcontact [kN]

100 P2

80

60

0 0.04 0.08 0.12


t [s]
30
20
aaxle [m/s2]

10
0
–10 0.04 0.08 0.12
–20 t [s]
–30

6
Mrail,dyn [kNm]

4
2
0
–2 0.04 0.08 0.12
–4 t [s]

12.8 Dynamic response of the wheel–rail system to an artificial rail


weld with both a long and short length-scale irregularity.
Rail welds 389

• the dynamic component of the displacement u of the rail/track, calculated


in a static co-ordinate system with its origin at the centre of the irregularity
and defined positive in downward direction;
• the wheel–rail contact force E for the wheel concerned (the dynamic
force is superimposed on the static value), in a convective reference
frame;
• the axle box acceleration a, in a convective reference frame;
• the dynamic component of the bending moment m in the rail, in a fixed
coordinate system at the centre of the irregularity.
From Fig. 12.8, a number of general observations can be made. The
rail (and this can to a large extent be generalised to the track) follows the
vertical irregularity quasi-instantaneously; this is clearly visible especially
for the short-length irregularity. The wheel displacement shows a delay in
response relative to the track. It does not show any influence of the short
peak in the irregularity, whereas this peak is reproduced almost exactly in
the rail displacements. The maximum dynamic contact force is determined
by the irregularity with the shortest length-scale. However, the corresponding
force peaks (P1 forces) are rather narrow, and the corresponding amount of
energy is rather small. These high-frequency peak forces damage mainly the
rail itself, as can be observed also from the dynamic rail bending moments.
The corresponding energy input will vanish by wave propagation in the
rail and dissipation in – mainly – the railpads. The highest energy of the
dynamic contact force is contained in the ‘carrier frequency’, which can be
clearly observed. This carrier frequency is related to the delayed reaction
of the wheel mass on the overall track stiffness to the irregularity (the P2
force). The relatively large energy contained in the relatively low carrier
frequency is mainly responsible for ballast bed deterioration, as it is not
efficiently dissipated in the rail and the railpads. Generally it can be stated
that components of irregularities having a longer length-scale than the wheel
radius or the sleeper span (0.5–0.6 m) damage the ballast-bed, whereas
components with shorter length-scales damage the railhead and rail (Fig.
12.9).
As has been mentioned, short components in a welding irregularity lead
to P1 forces, whereas longer components lead to P2 forces. This is mainly
due to the difference in inertia between the unsprung wheel mass and the

Increasing rail damage Increasing ballast bed deterioration

Rwheel dsleeper Length-scale


of rail welding
0 0.5 0.6 1m irregularities

12.9 Types of deterioration of the track system depending on the


length-scale of rail welding surface irregularities.
390 Wheel–rail interface handbook

equivalent track mass: assuming an equivalent track mass comprising two


fully effective concrete sleepers each of 300 kg and 4 m of effective rail mass
(60 kg/m; 60E1 profile) the equivalent track mass can be estimated at 850
kg. This is less than 50 % of the unsprung axle mass, which is approximately
1800 kg for conventional trains. The response of the wheelset to the short
excitation therefore shows a delay relative to the response of the track (rail
and sleepers). The P1 force, which is a quasi-instantaneous amplification of
the wheel-rail contact force, originates from the reaction of the rail and is
mainly determined by the rail properties (bending stiffness and inertia). The
P2 force results from the reaction of the wheelset on the track stiffness to
the excitation. Its magnitude is determined mainly by the unsprung wheel
mass and the equivalent track stiffness; the timescale of the peak is larger
due to the relatively low natural frequency of the dominating mass–spring
system.
Steenbergen and Esveld (2006b) investigated the quantitative relationship
between the actual rail weld geometry and the corresponding dynamic wheel–
rail response with the help of a FE model. Simulations were performed for a
sample of 239 rail weld measurements from the Dutch conventional network.
Model parameters were used as defined previously, except for the ballast
stiffness, which was taken as 78·106 N/m per sleeper, which is a conventional
value. The simulations were performed for different train velocities: 40,
80, 140 (conventional lines) and 300 km/h (high-speed lines) respectively.
For conventional lines, dynamic wheel–rail forces were generally below
70 kN. For a static nominal wheelload of 112.5 kN, this yields a dynamic
amplification factor of 1.6. This value is confirmed experimentally by Mutton
and Alvarez (2004). Figure 12.10 summarizes computational results from the
FE simulations. At the left, the relationship between the maximum dynamic
wheel-rail contact force and the maximum absolute first derivative (on 25
mm basis of a running five-point-average through the measurement signal,
which has a resolution of 5 mm) is shown for the different velocities for
which simulations were performed. For each velocity, a best linear fit through
the origin is displayed, disregarding any scatter.
Figure 12.10b shows the relationship between the contact force and the
train speed for a predefined maximum absolute first derivative, which is
taken as 5 mrad (again on 25 mm basis). In general, the relationship between
the force and the speed can be well approximated as linear. In terms of the
train speed V [m/s] and the maximum absolute rail inclination qmax [mrad]
(on 25 mm basis) the following expression is obtained (in [kN]):
Fdyn, max = x · V · qmax: x = 0.22 [12.1]
It is observed that x is a dimension-bearing coefficient. The issue has been
addressed by Steenbergen (2008a), showing that the validation coefficient
includes an equivalent track mass, a length dimension and a dynamic stiffness,
Rail welds 391

160

Max. dyn. contact force [kN]


140 v = 300 km/h
120
100
v = 140 km/h
80
60
v = 80 km/h
40
20
v = 40 km/h
0
0 2 4 6 8 10
Max. abs. inclination [mrad]

100
Max. dyn. contact force [kN]

Max. abs. inclination: 5 mrad


80

60
y = 0.28x
R2 = 0.97
40

20

0
0 40 80 120 160 200 240 280 320
Train speed [km/h]

12.10 Dependence of the dynamic wheel–rail contact forces at rail


welds on the train velocity, found from FEM simulations.

which is responsible for the scatter in the relationship. The linear relationship
between the maximum dynamic contact force (or the dynamic amplification
factor) and the train speed is confirmed experimentally by the results obtained
by Mutton and Alvarez (2004). A very similar result was found, both from
simulations and experiments, by Jenkins et al. (1974) at British Rail for the
peak forces occurring at dipped rail joints. This result was reflected in the
well-known Jenkins formula for the calculation of P1 and P2 forces. They
found these peak forces to be governed by the dip angle and the train speed;
furthermore the relationship between maximum dynamic contact forces and
both variables proved to be linear in a good approximation.
Jenkins’ results for rail joints, together with the results obtained for rail
welds, are shown in the graph in Fig. 12.11. In this figure, the dynamic peak
forces for rail joints and welds are depicted as a function of the product of
the maximum absolute inclination and the train speed; a ≈ tan a is defined
as the angle of a dipped rail end with the horizontal for rail joints, and as
the maximum absolute inclination (on 25 mm basis) of the geometry for rail
welds. It is clear that the force level for rail welds is significantly lower than
392 Wheel–rail interface handbook

350

Maximum dynamic wheel–rail force [kN] 300


P1 for rail joints

250

200
P2 for rail joints

150

100

50 Fdyn for rail welds

0
0 100 200 300 400 500
Tan a · v [mrad · m/s]

12.11 Maximum dynamic wheel–rail forces (Fdyn) for rail welds and
rail joints, as a function of the product of the maximum absolute
inclination (on 25 mm basis) or the dip angle and the speed.

for rail joints; the reduction factor is approximately three. This is partly due
to the quasi-static rail deflection under axle loading, which is much larger
for a jointed rail (behaving as a hinged beam) than for a welded rail. This
quasi-static deflection increases the dip angle at the moment of wheel passage
(Steenbergen, 2006).
Steenbergen (2008a,b) also investigated the dynamic wheel–rail interaction
at rail welds with a linear numerical–analytical frequency–domain model,
comprising a rail on elastic foundation and a train axle up to the primary
suspension level. This study was performed for a better understanding of
the backgrounds of wheel–rail interaction at short irregularities and trend
behaviour. Analytical results show a perfectly linear relationship between
the slope of a ramp in the rail geometry and the magnitude of the dynamic
contact force in the time domain. The linear trend of this contact force as
a function also of the train velocity is shown in Fig. 12.12, for the case of
a ramp of 2 mrad and a length of 25 mm. Thus, for the elementary case
of a ramp in the rail surface, the first peak force (P1) may be considered
as directly proportional to the speed and the ramp angle, according to the
expression:
Fdyn,max = x · V · qmax: x = 0.18 [12.2]
Rail welds 393

35

Maximum dynamic wheel–rail force (P1) [kN]


30

25

20
Fit ---:
15 y = 0.36 x;
R2 = 1
10

0
0 15 30 45 60 75 90
Velocity [m/s]

0 40 80 120 160 200 240 280 320


Velocity [km/h]

12.12 Relationship between the maximum dynamic contact force


(P1) and the velocity, for a given ramp inclination (2 mrad) and basis
length (25 mm).

In the analytical simulations, a running five-point average of the measurement


signal of a weld has been used as calculation input, with a discretization
basis of 5 mm. The graphs in Fig. 12.13 show examples of the computed
time histories of the contact force for two measured welds, for a conventional
passenger train velocity of 140 km/h. The first weld is rather well aligned,
but shows the effect of non-homogeneous shrinkage of the weld and parent
material after the grinding process, which has obviously been performed at
too high a temperature. The second weld is smooth on a micro-scale, but
shows a severe misalignment before the rail ends have been welded together,
resulting in a pronounced step of about 2 mm in rail height over 25 mm. For
this particular weld, contact loss will occur as soon as the dynamic force
component exceeds the static pre-load.
In Fig. 12.14 an example is shown of the correlation between the calculated
maximum absolute dynamic wheel–rail contact forces (during the time interval
of overpassing the weld) and the vertical absolute peak deviations (Fig.
12.14a) or the maximum absolute inclinations of the welds (Fig. 12.14b),
for a train velocity of 300 km/h. Similar to the results obtained earlier with
the FE model, the correlation between forces and maximum inclinations is
found to increase significantly with the velocity. This can be explained by
a more flat transfer of the system as the excitation frequency band shifts
towards higher frequencies with increasing speed.
394 Wheel–rail interface handbook

1.2
Measured geometry
z [mm]

0.4 5-point smoothing

–0.4 –0.4 –0.3 –0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5
x[m]
–1.2
350

250

150
Fdyn [kN]

50

–50 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
   Vt[m]
–150 Static wheel load
(a)

0.04
z [mm]

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
–0.04    x[m]
Measured geometry
20 5-point smoothing

10
Fdyn [kN]

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
      Vt[m]
–10

–20
(b)
12.13 Examples of measured welds, their geometry after 5-point
smoothing and calculated time histories of the dynamic wheel–rail
contact force at 140 km/h.

The linear relationships, obtained from the numerical–analytical


computations, between the dynamic force and the maximum absolute
inclination are summarized in Fig. 12.15a for all train velocities. Figure
12.15b shows the dependence of the dynamic force on the train speed for a
given maximum inclination (5 mrad, on 5 mm basis of the five-point averaged
measurement signal). Again, the relationship between the dynamic force and
the train speed can be very well described as directly proportional.
Figure 12.15 yields the value x = 0.19 for the coefficient in the general
relationship (Eq. 12.1). From the FE analysis x = 0.22 was found. This
Rail welds 395

200
y = 152.56x
v = 300 km/h
R2 = 0.18

Max. abs. dan. contact force [kN]


160

120

80
y = 100.45x + 28.12
R2 = 0.35
40

0
0 0.3 0.6 0.9 1.2 1.5
Vertical abs. peak deviation [mm] (unsmoothed)
(a)
200
v = 300 km/h y = 15.82x
R2 = 0.86
160

120

80

40

0
0 4 8 129 16
Max. abs. inclination [mrad) (5-point smoothed)
(b)
12.14 Peak value of the dynamic wheel–rail force as a function of the
vertical peak deviation (a) and the maximum absolute inclination (b)
respectively, at 300 km/h.

slightly higher value can be explained from the fact that inclinations on a 25
mm basis have been used in FEM analysis. The inclination decreases when
the sample interval is extended from 5 to 25 mm. It can be concluded that
x ≈ 0.2 in practice (for 54 E1 rail).

12.5 Rail weld geometry assessment; the Dutch rail


welding regulations (2005)
Traditionally, the vertical geometry of rail welds is assessed manually with
a steel straightedge in combination with a feeler gauge, using the principle
396 Wheel–rail interface handbook

250
300 km/h

Max. abs. dyn. contact force [kN] 200

200 km/h
150

140 km/h
100

80 km/h
50

40 km/h
0
0 4 8 12 16
Max. abs. inclination [mrad] (5-point smoothed)
(a)

100
Max. abs. dyn. contact force [kN]

Inclination : 5 mrad

y = 0.94x
75 R2 = 0.99

50

25

0
0 10 20 30 40 50 60 70 80 90
Train speed [m/s]
(b)

12.15 Dependence of the dynamic wheel–rail contact force at welds


on the maximum absolute inclination (for different train speeds) (a)
and on the train speed (for 5 mrad, 5 mm basis) (b).

of tolerances. These tolerances are defined for a given basis length, which
is commonly 1 m. A common value for the vertical tolerance is 0–0.3
mm (Esveld, 2001). Two drawbacks of this traditional method are that no
restrictions are being posed to the ‘smoothness’ of the rail surface, whereas
this shape has a direct relation to the magnitude and the spectrum of the
dynamic wheel–rail interaction force, and the fact that the train speed has
no influence, whereas its effect on the contact forces is far from negligible.
The spectrum of the wheel–rail forces in its turn is related to the rate of track
degradation. A typical result of the current rail welding assessment methods
(especially aluminothermic in-situ welding) is shown in Fig. 12.16. After the
Rail welds 397

0.6
y [mm] ‘Set-up’ Grinding area ‘Set-up’
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1
x [m]

12.16 Typical example of the vertical geometry along an in-situ made


rail weld, resulting from traditional assessment methods.

welding process, the ‘top’ resulting from the aligning of both rail ends, as far
as it does not fit within the tolerance, is ground off. The result is often far
from smooth, and introduces large dynamic variations in contact stresses.
the development of digital straightedges that sample the vertical rail
geometry enables more advanced assessment methods, that do account
for the ‘smoothness’ of the rail geometry. it is not feasible to compute the
wheel–rail contact forces in each step of weld grinding, in order to limit the
dynamic contact stresses. A compromise between theoretical efficiency and
practical feasibility, at the cost of exactitude, can be looked for by correlating
a parameter describing the geometry of the weld and the maximum contact
force that would occur for a predefined reference track and wheelset, and
the train speed. the relationship derived earlier (eq. 12.1) could be used
for this purpose, accounting for both the ‘smoothness’ of the rail surface
through its maximum first derivative and for the train speed. For assessment
purposes, the dimensionless rail weld quality index (QI) is proposed. It is
defined as the actually measured maximum absolute first derivative (on 25
mm basis) normalized with a norm value, which can be defined differently
for different values of the line section speed:
dz i (x )
dx max, actual
QI = (≤ 1: acceptance; >1: rejection) [12.3]
dz
dx norm
nor

the norm values for the inclinations dz /d/dx norm


nor for different line section
speeds must then be established such that the dynamic force, according to
eq. (12.1), is constant. Line section speed intervals are chosen according to
typical values: 40, 80, 140, 200 and 300 km/h. Typical examples of welds
for which a more ‘dedicated’ weld assessment would make sense are shown
in Figs 12.17–12.19 (Steenbergen and esveld, 2006a).
Figure 12.17 shows an almost perfectly straight weld; however, showing
small indentations due to non-uniform shrinkage after welding and hot
grinding. the weld in Fig. 12.18 shows an aggressive step due to a bad
alignment of the rail ends before welding. this is accepted according to
398 Wheel–rail interface handbook

0.2
y [mm]

0
–0.2
0 0.2 0.4 0.6 0.8 1
x [m]

12.17 Example of a weld with indentations due to non-uniform


shrinkage after welding and grinding.

0.3
y [mm]

0.1
–0.1
0 0.2 0.4 0.6 0.8 1
x [m]

12.18 Example of a weld with a pronounced step due to bad rail end
alignment.

0.2
y [mm]

0
–0.2
0 0.2 0.4 0.6 0.8 1
x [m]
12.19 Example of a weld with a ‘smooth’ surface but with negative
height coordinates.

many standards allowing a vertical tolerance of +0.3 mm. However, owing


to the large magnitude of the first derivative at the transition between both
rail ends the weld would be rejected. In Fig. 12.19, an example is shown
of an almost perfect weld. However, because it has some negative height
co-ordinates, it is rejected according to traditional standards, which do not
allow negative values. The derivatives do not show pronounced peaks and
therefore the weld would be accepted.
The norm value for high-speed lines (300 km/h), would be the strictest
one in the speed range. Its value can be based on the geometrical quality of
new straightened rail: there is no reason why rail welds should have a better
geometrical quality than the rail itself. This quality has been determined by
measurements on new straightened longrails (120 m); a sample population
of 100 segments of 1 m length of newly laid rails on the Dutch high-speed
line HSL-South have been measured (Esveld, 2005). The distribution function
of the absolute maximum first derivatives (25 mm basis) for these segments
is shown in Fig. 12.20. The 95 percentile value of the first derivative in this
figure is 0.7 mrad. Therefore, 0.7 mrad could be taken as an appropriate value
for the intervention level of first derivative of the weld geometry for high-
speed lines. This accuracy is very close to the maximum obtainable accuracy
in welding and grinding with grinding trains. Tests in this regard have been
Rail welds 399

1
95%

0.8
Cumulative frequency

0.6

CDF of weld quality on the HSL-South


0.4 test section, West, after Speno final grinding
test section, East, after Speno final grinding
296 welds, after GWM grinding
0.2
100 measurements on new straightened rail

0.7 1.1 1.3


0
0 0.5 1 1.5 2 2.5 3
Max. abs. rail inclination [mrad]

12.20 Cumulative distributions of maximum absolute first


derivatives (25 mm basis) on the Dutch HSL-South; comparison
of measurements on new straightened rail segments and
measurements of rail welds after grinding with grinding trains.

carried out during the installation of the Dutch HSL-South (Winter et al.,
2007). In Fig. 12.20, the cumulative distribution of the maximum absolute
first derivatives (25 mm basis) of 296 test welds (a mix of flash-butt and
thermite welds) is also shown before and after grinding with a GWM 550
grinding train. Before grinding with the GWM the 95 percentile value of the
maximum first derivative is 5.3; this reduces to 1.1 afterwards. Finally, Fig.
12.20 displays the cumulative distributions of the maximum absolute first
derivative of the rail welds on a double track test section (East and West)
of 6 km on the HSL-South. These welds were manually pre-ground, and
finally ground with a Speno grinding train (types RR 24 and 48). Based on
the results in Fig. 12.20, the 0.7 mrad limit seems too strict a value, and 1
mrad is proposed as a limit value in the QI determination for welds in new
high-speed tracks (with norm speed 300 km/h).
In order to consistently determine speed-dependent norm values for the
maximum inclination, a pre-defined dynamic force level should be adopted.
This force level should be related to the damage level but, because little
is known of this relationship and furthermore this is strongly frequency-
dependent, any choice of this level is rather arbitrary. Two facts may be
taken as a point of departure:
1. the empirical vertical tolerance of 0.3 mm has been used worldwide for
several decades and has not led to systematic problems; the maximum
peak deviation resulting from the standardized inclination should, for
140 km/h (which is a conventional passenger train speed), not have a
different order of magnitude;
400 Wheel–rail interface handbook

2. the norm value for high-speed lines (300 km/h) has been established at
the feasibility limit of 1 mrad.
The norm value of 1 mrad at 300 km/h leads, according to Eq. (12.1)
(with x = 0.2), to a dynamic force level of 17 kN. The value of 0.7 mrad,
obtained for new straightened longrail, leads to a force level of 12 kN. The
curves according to Eq. (12.1) corresponding with both force levels are
depicted in Fig. 12.21. The following speed intervals have been used: 0–40
km/h, 40–80 km/h, 80–140 km/h, 140–200 km/h (conventional lines) and
200–300 km/h (high-speed lines).
The norm values should ideally be situated in the hatched area between
the two curves in Fig. 12.21, with a cut-off by the feasibility limit. However,
due to the hyperbolical behaviour, the values for low speeds grow to infinity,
whereas it is easy to grind, even manually, to a much better quality. Therefore,
the smooth function depicted in Fig. 12.21, leaving the hatched area at 40
and 80 km/h, has been adopted by the Dutch Rail Infra Manager ProRail.
In 2005, the geometrical standards as shown in Table 12.1 were introduced
for metallurgical rail welds in the Netherlands.
Table 12.2 gives the amplitudes of a sinusoidally shaped weld with a half
wavelength of 1 m, and corresponding to the inclinations in Table 12.1. These
amplitudes give an impression of the maximum vertical peak deviations
corresponding to the inclination values. Using the 1.8 mrad criterion for 140
km/h, the maximum amplitude equals 0.57 mm, which is larger than – but
in the same order of magnitude as – the 0.3 mm tolerance (1).

8
Inclination norm value (25 mm basis) [mrad]

Based on 1 mrad at 300 km/h (17 kN)


7 Based on 0.7 mrad at 300 km/h (12 kN)
Adopted by ProRail
6

2
Feasibility limit
1

0
0 40 80 120 160 200 240 280 320
Train speed [km/h]

12.21 Determination of speed-dependent norm values in the QI


definition.
Rail welds 401

Table 12.1 Inclination norm values adopted by ProRail for the QI


determination (ProRail, 2007)

Train speed [km/h] Max. absolute inclination


(25 mm basis) [mrad]

0–40 3.2
40–80 2.4
80–140 1.8
140–200 1.3
200–300 1
NB: 60, 100 (special tracks) 2.8, 2.2

Table 12.2 Inclination norm values (25 mm) and the resulting 2 m wave
amplitude

Train speed [km/h] Inclination values Resulting 2 m wave


[mrad] amplitude [mm]

  40 3.2 1.02
  80 2.4 0.76
140 1.8 0.57
200 1.3 0.41
300 1 0.32

Norm values
40 km/h
300
200
140
80

0.8
73 %
Cumulative frequency

0.6
58 %
46 %
0.4 CDF of 239 weld measurements

27 %
0.2 16 % satisfies tolerance 0–0.3 mm
17 %

0
0 1 2 3 4 5 6 7 8 9 10
Max. abs. weld inclination (25 mm basis) [mrad]
12.22 Comparison of acceptance levels of rail welds according to the
traditional method based on tolerances and the QI-method.

In Fig. 12.22, a comparison is made of the acceptation level of welds,


according to the conventional standards based on vertical tolerances (0–0.3
mm, independent of the speed), and the method based on first derivatives.
The analysis has been performed for the sample of 239 welds which has
402 Wheel–rail interface handbook

been used for the computations earlier in this chapter. It is observed that
the acceptance level increases drastically (except for 300 km/h). At the
conventional 140 km/h the acceptance increases by 30 %. This relaxation,
which, according to the investigation in this chapter should not lead to an
increasing rate of deterioration, is due to the fact that peak deviations larger
than 0.3 mm are accepted, provided that the contact geometry is smooth.
The weld geometry assessment method elaborated in this section can be
simply implemented in practice. This is briefly illustrated in the following.
After welding and cooling down of the rails, the quality of the geometry
can be measured using an electronic digital straightedge. In the processor
the routines for the calculation of the quality index can be programmed. The
device samples the rail geometry and plots the normalized first derivative
(dependent on the train speed) and the QI on a screen. With the help of
this output, the geometry can be optimised such that the standards are met.
An example of such a screen output is given in Fig. 12.23. At the location
where the longitudinal rail geometry shows a relatively large inclination,
the requirement is not met (the QI is 1.7), and the weld should be ground
before acceptance. In Fig. 12.24, an example is shown of rail weld geometry
assessment using an electronical straightedge.

12.6 Welding irregularities, energy considerations


and deterioration
Previously, an assessment method of welding irregularities based on a limitation
of dynamic forces in the wheel–rail interface was described. However, in
many cases deterioration of a track component is directly related to energy

mm 2 2
QI: 1.7
v: 40 km/h
1.5

1 1

0.5

0 0
0.2 0.4 0.6 0.8 1

–1
Weld geometry Normalised 1st derivative Norm – per definition 1

12.23 Example of the screen output of a digital measuring device for


the assessment of the geometry of rail welds.
Rail welds 403

(a)

(b)

12.24 Ground rail weld (a) and geometry assessment (b) using a
digital straightedge.

dissipation. Damage occurrence can then be translated into reaching a saturation


limit in energy dissipation capacity in time. Therefore, in future regulations
it would seem more appropriate to limit the power spectrum corresponding
404 Wheel–rail interface handbook

to a given track geometry in the whole wavelength or frequency domain.


Although the step from purely geometrical to force-based track assessment
is being made, the step from force-based to input-power-based assessment
is not yet feasible. However, some examples focusing on energy-based
assessment are given in this section for illustration.
Basically, a correlation between the maximum first derivative (or another
geometrical parameter describing the weld geometry) and the energy or
maximum power input into the track can be sought. This correlation is
different from that between the maximum first derivatives and the maximum
contact forces, due to the fact that the dynamic stiffness spectrum of the
track (including effects from both elasticity and inertia) is a non-uniform
spectrum. Depending on the geometry of the excitation, different frequency
regions of the dynamic track stiffness spectrum play a role.
Figure 12.25 shows the contact force, the vertical rail velocity and the
power input along the track (or energy input into the track section) for both
welds displayed in Fig. 12.13, when a train axle passes at 140 km/h. The
first weld, which is rather smooth but suffers from some shrinkage of the
weld material, yields a maximum power input only between 2 and 3 kW.
The second weld, containing a severe step, yields a maximum power input

0.08
0.04
z [mm]

0
0.2 0.4 0.6 0.8 1
–0.04
x = Vt [m]
–0.08
20
Dyn. force [kN]

10

0
0.2 0.4 0.6 0.8 1
–10   x [m]

–20
3

2
Power [kW]

Einput
1

0
0.2 0.4 0.6 0.8 1
–1   x [m]
(a)

12.25 Power input into the track per train wheel for the welds
depicted in Fig. 12.13 at 140 km/h.
Rail welds 405

1.2

0.4
z [mm]

–0.4 0.2 0.4 0.6 0.8 1


x = Vt [m]
–1.2
350
Dyn. force [kN]

250
150
50
–50 0.2 0.4 0.6 0.8 1
–150   x [m]

800
Power [kW]

600

400 Einput

200

0
0.2 0.4 0.6 0.8 1
  x [m]
(b)

12.25 Cont’d

between 700 and 800 kW, for each passing wheel (using a linear calculation).
Given that modern locomotives generally have a traction power between
2000 and 8000 kW (the latter for high-speed trains), a power loss of 800
kW per wheel into the track due to a bad weld is very severe.
Figure 12.26 shows computational results for the elementary case of a
small ramp in the rail surface, which was considered in Fig. 12.12. In Fig.
12.26a, maxima of the power input into the track due to P1 are shown as
a function of the ramp inclination, for a given train speed of 140 km/h.
In Fig. 12.26b this is repeated as a function of the velocity, for a given
inclination of 2 mrad. The basis length of the ramp equals 25 mm in both
cases. In both cases, the relationship is exactly second-order polynomial.
Both relationships with the maximum dynamic force were linear. This indicates
the severity of rail surface irregularities for long-term track behaviour, which
is governed by energy dissipation mechanisms. These mechanisms need
future investigation.
406 Wheel–rail interface handbook

50
v = 140 km/h

Maximum power input due to


40 v = 0.33x2 + 0.04x
R2 = 1,00

30
P1 [kW]

20

10

0
0 2 4 6 8 10 12
Inclination [mrad]

7
q = 2 mrad
Maximum power input due to

6
y = 7.86e–05x2 –1.59E–03x
5 R2 = 1.00E + 00
P1 [kW]

0
0 40 80 120 160 200 240 280 320
Velocity [km/h]

12.26 Maxima of the power input into the track due to P1 for a
ramp with a basis of 25 mm. (a) Values as a function of the ramp
inclination, for 140 km/h; (b) values as a function of the velocity, for
an inclination of 2 mrad.

12.7 References
Böhmer A and Klimpel T (2002), Plastic deformation of corrugated rails – a numerical
approach using material data of rail steel, Wear, 253, 150–61.
Cai W, Wen Z, Jin X and Zhai W (2007), Dynamic stress analysis of rail joint with
height difference defect using finite element method, Engineering Failure Analysis,
14, 1488–499.
Chen Y, Lawrence F V, Barkan C P L and Dantzig J A (2006a), Heat transfer modelling
of rail thermite welding, Proc IMechE, Part F: Journal of Rail and Rapid Transit,
220, 207–17.
Chen Y, Lawrence F V, Barkan C P L and Dantzig J A (2006b), Weld defect formation
in rail thermite welds, Proc IMechE, Part F: Journal of Rail and Rapid Transit, 220,
373–84.
Rail welds 407

Esveld C (2001), Modern Railway Track (2nd edn), Zaltbommel, the Netherlands, MRT-
productions.
Esveld C (2005), Measurements of Rail and Weld Geometry on HSL-South, ECS report,
Zaltbommel, the Netherlands.
Hiensch M, Nielsen J C O and Verheijen E (2002), Rail corrugation in the Netherlands
– measurements and simulations, Wear, 253, 140–49.
Ilic N, Jovanovic M T, Todorovic M, Trtanj M and Saponjic P (1999), Microstructural and
mechanical characterization of postweld heat-treated thermite weld in rails, Materials
Characterisation, 43, 243–50.
Ishida M, Moto T, Kono A and Jin Y (1999), Influence of loose sleeper on track dynamics
and bending fatigue of rail welds, Quarterly Report of RTRI, 40, 80–85.
Jenkins H H, Stephenson J, Clayton G A, Morland G W and Lyon D (1974), The effect
of track and vehicle parameters on wheel/rail vertical dynamic forces, Railway
Engineering Journal, January, 2–16.
Kabo E, Nielsen J C O and Ekberg A (2006), Prediction of dynamic train-track interaction
and subsequent material deterioration in the presence of insulated rail joints, Vehicle
System Dynamics, 44, 718–29.
Li Z, Zhao X, Esveld C and Dollevoet R (2006), Causes of squats: correlation analysis
and numerical modeling, Proceedings 7th International Conference on Contact
Mechanics and Wear of Rail/Wheel Systems, Brisbane, Qld Australia, 24–27 September,
439–46.
Li Z, Zhao X, Esveld C and Dollevoet R (2007), Rail stresses, strain and fatigue under
dynamic wheel-rail interaction, Proceedings International Heavy Haul Association
Specialist Technical Session, Kiruna, Sweden, 11–13 June, 389–96.
Mutton P J (2000), Material aspects of weld behaviour in wheel-rail contact, Proceedings
Fifth International Conference on Contact Mechanics and Wear of Rail/Wheel Systems,
Tokyo, Japan, 25–27 July, 131–5.
Mutton P J and Alvarez E F (2004), Failure modes in aluminothermic rail welds under
high axle load conditions, Engineering Failure Analysis, 11, 151–66.
ProRail (2007), Directives RLN00127–1&2, Part 1 – Operationele eisen voor metallurgische
lassen in bovenbouwconstructies; Part 2 – Kadereisen voor metallurgische lassen in
bovenbouwconstructies, ProRail, Utrecht, the Netherlands.
Shitara H, Terashita Y, Tatsumi M and Fukada Y (2003), Nondestructive testing and
evaluation methods for rail welds in Japan, Quarterly Report of RTRI, 44, 53–8.
Skyttebol A (2004), Continuous Welded Railway Rails: Residual Stress Analyses, Fatigue
Assessments and Experiments, PhD Thesis, Chalmers University of Technology,
Gothemburg, Sweden.
Skyttebol A, Josefson B L and Ringsberg J W (2005), Fatigue crack growth in a welded
rail under the influence of residual stresses, Engineering Fracture Mechanics, 72,
271–85.
Steenbergen M J M M (2006), Modelling of wheels and rail discontinuities in dynamic
wheel-rail contact analysis, Vehicle System Dynamics, 44, 763–87.
Steenbergen M J M M (2008a), Quantification of dynamic wheel-rail contact forces at
short rail irregularities and application to measured rail welds, Journal of Sound and
Vibration, 312, 606–29.
Steenbergen M J M M (2008b), Wheel–Rail Interaction at Short-Wave Irregularities,
PhD Thesis, Delft University of Technology, Delft, the Netherlands.
Steenbergen M J M M and Esveld C (2006a), ‘Rail weld geometry and assessment
concepts’, Proc. IMechE, Part F: Journal of Rail and Rapid Transit, 220, 257–71.
408 Wheel–rail interface handbook

Steenbergen M J M M and Esveld C (2006b), Relation between the geometry of rail welds
and the dynamic wheel–rail response; numerical simulations for measured welds, Proc
IMechE, Part F: Journal of Rail and Rapid Transit, 220, 409–24.
Tawfik D, Kirstein O, Mutton P J and Chiu W K (2006), Verification of residual stresses
in flash-butt-weld rails using neutron diffraction, Physica B, 385–6, 894–6.
Tawfik D, Mutton P J and Chiu W K (2008), Experimental and numerical investigations:
alleviating tensile residual stresses in flash-butt welds by localised rapid post-weld
heat treatment, Journal of Materials Processing Technology, 196, 279–91.
Terashita Y and Tatsumi M (2003), Analysis of damaged rail weld, Quarterly Report of
Railway Technical Research Institute, 44, 59–64.
Webster P J, Mills G, Wang X D, Kang W P and Holden T M (1997), Residual stresses
in alumino-thermic welded rails, Proc. IMechE, Journal of Strain Analysis, 32,
389–400.
Wimmer W E, Connell D A and Boos M J (2002), Joint elimination through mobile flash-
butt welding on Union Pacific Railroad, Proc. of Innotrans 2002, Berlin, Germany,
24–27 September.
Winter T, Meijvis P A J, Paans W J M, Steenbergen M J M M and Esveld C (2007),
Track quality achieved on HSL-South – reduction of short-wave irregularities cuts
life cycle cost, European Railway Review, 13(3), 48–53.
Wöhnhart A and Wenty R (2002), Mobile flash-butt welding: three decades of experience,
Rail Engineering International, 3, 11–16.

You might also like