Sei sulla pagina 1di 13

Engineering Failure Analysis 45 (2014) 151–163

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

An investigation of distortion-induced fatigue cracking under


variable amplitude loading using 3D crack propagation analysis
Mustafa Aygül a,⇑, Mohammad Al-Emrani a, Zuheir Barsoum b,c, John Leander b
a
Chalmers University of Technology, Gothenburg, Sweden
b
Royal Institute of Technology, Stockholm, Sweden
c
Khalifa University of Science Research and Technology, Abu Dhabi, United Arab Emirates

a r t i c l e i n f o a b s t r a c t

Article history: The distortion-induced fatigue cracks in the welded details of the Söderström Bridge are
Received 19 December 2013 analytically and numerically investigated by performing 3D crack propagation analyses
Received in revised form 11 May 2014 with variable amplitude fatigue loading. In the crack propagation analyses, the effects of
Accepted 20 May 2014
bridge loading are defined on the basis of the field measurements in order to simulate crack
Available online 18 June 2014
growth and predict the residual fatigue life of the studied detail as accurately as possible.
The effect of crack closure and crack direction while considering the most common criteria
Keywords:
is also studied. The results are compared with those obtained from the crack propagation
Distortion-induced fatigue cracking
Out-of-plane distortion
analyses with constant amplitude fatigue loading presented in Aygül et al. [1].
Variable amplitude loading The results of the crack growth simulations with variable amplitude fatigue loading have
Crack propagation analysis generally shown good agreement with the real crack formation and reveal that the crack
Mixed-mode conditions growth rates are different in different directions. The crack behaviour in the damaged
detail is mainly controlled by the loading and geometrical arrangement of the detail com-
ponents. There is generally a significant difference between constant and variable ampli-
tude fatigue crack growth analyses and the variable amplitude fatigue crack growth
analyses yield more conservative results. The main reason for this difference is the bridge
loading and the number of stress cycles defined in the analyses. The crack direction criteria
studied in this investigation showed basically the same crack formation and crack growth
rate.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction

In large welded structures such as railway bridges, the estimation of the fatigue design life is usually performed using the
linear cumulative damage method known as the Palmgren–Miner rule or the equivalent damage method. The common pro-
cedure when using these methods is that the load effects on the structural components caused by the variable amplitude
fatigue load need to be represented by one or more equivalent constant amplitude stress ranges which are easier to manage
in a design situation [2–4]. It is assumed theoretically that the equivalent constant amplitude stress range results in approx-
imately the same fatigue damage as it does in variable amplitude loading. This is also due to the fact that the fatigue strength
curves recommended by design standards were obtained from fatigue tests under constant amplitude (CA) cycling loading,

⇑ Corresponding author. Tel.: +46 31 772 22 49; fax: +46 31 772 22 60.
E-mail address: mustafa.aygul@gmail.com (M. Aygül).

http://dx.doi.org/10.1016/j.engfailanal.2014.05.015
1350-6307/Ó 2014 Elsevier Ltd. All rights reserved.
152 M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163

i.e. S–N curves based on constant stress ranges. The equivalent stress range or transformed set of representative constant
amplitude stress ranges are then used to perform fatigue design or analysis either directly by applying the Palmgren–Miner
damage accumulation rule, or by using the equivalent constant amplitude stress range concept [5]. These procedures are rec-
ommended in various fatigue design standards such as EC3 [6], BS5400 [2] and AASHTO [4].
The estimation of the residual service life of damaged bridge details is, however, based on the fracture mechanics methods
which provide information about crack size and anticipated crack growth rates. For fatigue assessments of welded structures,
one common method is linear-elastic fracture mechanics. When a crack is detected in a structure in service, the most reliable
way accurately to represent the actual traffic load is to record sufficient stress histories directly from the structure. Fatigue
evaluation based on field-measured stress-range histograms is assumed to be a more accurate and efficient method for eval-
uating existing bridges [5]. The stress histogram, including the constant amplitude stress ranges with corresponding cycles
from the measured stress history, can be created using a counting method such as the Rainflow method, in the same way
as in the design phase. This involves the load effects being defined more accurately when considering the real traffic load
to estimate the residual life. However, the representative constant amplitude stresses do not include the physical significance
of bridge loads, i.e. compressive stresses and overloads, as well as underloads. On the other hand the above-mentioned stress
counting method effectively accommodates small intermediate stress ranges and the coupling of larger maximum and lower
minimum stresses in comparison with other stress counting methods, such as peak count methods and reservoir count meth-
ods [7]. These factors may have a decisive impact on crack behaviour, i.e. crack size and crack growth rates, during the service
life of structures.
Distortion-induced fatigue cracking is without question a common problem in steel and composite girder bridges and the
residual life assessment of this type of cracking is somewhat complex, due to secondary stresses which play a significant role
in crack behaviour. The accuracy of the life assessment methods might therefore be dependent not only on the stress ranges
generated by variable amplitudes but also on other parameters that might affect the fatigue performance of damaged bridge
details, such as the mean stress, load sequence, underload and overload.
In this paper, the behaviour of distortion-induced fatigue cracks developed in the web gaps in a multi-girder welded steel
bridge, the Söderström Bridge, located in Stockholm, was studied by performing 3D crack propagation analysis under vari-
able fatigue loading in amplitudes based on linear-elastic fracture mechanics (LEFM), including the effect of both fracture
modes and crack closure. The main purpose of this study is to examine the behaviour of distortion-induced fatigue cracks
under more realistic fatigue loading conditions by considering load history-induced sequence effects and to compare the
accuracy of the results from various crack propagation analyses.

2. Distortion-induced fatigue cracking and VA fatigue loading

Distortion-induced fatigue cracking is the most common damage problem in steel and composite girder bridges and it has
been discussed in more detail in [1]. This type of fatigue damage is mainly generated by out-of-plane deformation in web
regions where transverse cross-beams are connected to the web of the longitudinal main girder, as shown in Fig. 1. This
out-of-plane deformation can induce very high local stresses, even though the deformation is relatively small in comparison
with the deformation in structural components. It is therefore important to represent these small deformations as accurately
as possible in any investigations.
Load effects generated by traffic on steel and composite bridges are generally complex. A load history is generally
composed of stress cycles which vary in amplitude, containing both compressive and tensile stresses, as shown in Fig. 2.
In conventional fatigue design and because welded details are always assumed to have residual tensile stresses, fatigue cal-
culations are made while considering the stress range irrespective of stress ratio. Simplifying the computation of the equiv-
alent constant amplitude stress range may involve ignoring some effects which might be important with respect to the
fatigue strength of the studied detail. It is well known that the strength of a damaged material depends on the size of cracks
and the stresses present in the material. For distortion-induced fatigue cracking, the stresses at the crack location are caused

Fig. 1. Out-of-plane distortion in a web plate of a main girder and bending moments [1].
M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163 153

Fig. 2. The measured load history of the studied detail.

by the secondary effects and, when these effects are reduced due to any repairs or relaxation, the stresses causing a crack to
initiate and propagate will probably be reduced and with them the crack growth.
The main purpose of this study is to examine the propagation of fatigue cracks generated by distortional effects under
variable amplitude loading. Fatigue cracks of this kind account for the majority of fatigue damage reported for steel and com-
posite bridges and there is a clear need for a reliable methodology to estimate the residual life of the welded bridge details
with this kind of cracking [8–13]. In this study, crack propagation analysis with the finite element method has been
employed, assisted by field measurements (the FE models and the field measurements are presented in detail in [1]). Crack
propagation analyses were then carried out to study the crack behaviour using the real traffic volume and loads based on the
monitored data. Since the crack was non-planar and influenced by the secondary stresses, mode-I and mixed-mode condi-
tions were also considered in the crack propagation analyses, even though mode I is the governing mode condition for the
studied detail, as stated in [1].

3. Case study

The studied steel railway bridge, the Söderström Bridge, is located in Stockholm and the cracked welded details (shown in
Fig. 3) in this bridge were presented in [1]. The global FE model whose accuracy was verified by the measured data and the FE
models used for the crack propagation analyses were also presented in the same reference. In order to compare the results
obtained from Ref. [1] with those obtained in this study, the same FE models were used and these models are therefore not
presented in this paper.

3.1. Crack propagation analysis

In residual fatigue life evaluation, the most important factor is accurately to define the load-induced stress ranges which
might be obtained by the field strain measurements. On the Söderström Bridge, the strain measurements using a total of 54

Fig. 3. The Söderström Bridge and the crack location detected in the web of the main girder.
154 M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163

strain gauges were continuously collected in different locations on the bridge under daily traffic loads for two months [14].
The stress ranges defined for the crack propagation analyses with variable amplitude fatigue loading are based on this field
measurement located 10 mm from the weld toe on the studied welded detail, corresponding to one week’s railway traffic
load, see Figs. 2 and 4. As seen in this figure, the traffic load generates both tensile and compressive stresses in the welded
region which, from a fracture mechanics point of view, means that the crack region is loaded by the stresses causing crack
growth rate acceleration and retardation.
The load history shown in Fig. 2 is defined in the crack propagation analysis by using the X60 train type in the global
model. The load position of the X60 train type was determined by considering the maximum stress in the vertical direction,
which was the measurement direction. It was also shown that, in this position, the maximum out-of-plane distortion in the
cracked region was obtained from the FE analysis. This maximum vertical stress was the reference stress and it was used to
scale down in order to obtain the stresses shown in Fig. 2. In this way, the measured load history was applied to proportional
fatigue loading obtained from the X60 train type in the crack propagation analyses. A part of the global model and the
location of the defined stress in the studied welded region are shown in Fig. 4.
In the crack propagation analyses, the size, shape and location of the initial crack, as well as the crack increment, were
chosen to be the same as in [1]. A local model containing the initial crack was introduced and used in the crack growth anal-
yses. In the crack growth simulations, Abaqus v6.11 software was used as a solver in order to calculate the displacements in
the updated model, while Franc3D v6.0 was used to calculate the new crack front based on the displacements obtained from
Abaqus v6.11 and for re-meshing the local model for the extended crack.

3.2. Crack closure effects

Fatigue crack growth is a complex process influenced by many variables such as loading, geometry, material microstruc-
ture and environmental factors, all of which might have an important effect on the crack growth rate. In cycling load con-
ditions, crack closure may occur when the crack surfaces are in contact, especially at low stress ratios, causing smaller crack
tip opening displacements, i.e. the crack is not fully open during the entire load cycle. Prior to the research reported by Elber
[15], it was assumed that crack closure was only caused by the loading parts that produce compressive stresses. However,
Elber [15] showed that crack closure occurs even underloads that produce tensile stress at the crack front and that this
mechanism has significant effects on the acceleration and retardation of fatigue crack propagation. According to Elber, a
crack closure exists during a lower part of the load cycle, Kop (the stress level to open crack surfaces), since the crack remains
closed. The reason for crack closure under low tensile stress is the effect of plasticity left in the wake of the growing crack,
since the stress at the crack tip is significantly higher, causing plasticity in the crack tip region. When the crack surfaces come
into contact, there is no change in crack-tip strain during the cyclic loading of a closed crack [15]. The cyclic load which pro-
duces compressive stress and the lower portion of cyclic loading generating tensile stress do not therefore contribute to
crack propagation. Crack closure reduces the fatigue crack growth rate by reducing the stress intensity at the crack tip.
It has been reported that lower R ratios contribute to lower crack growth rates, while higher R ratios contribute to higher
crack growth rates, i.e. mean stress sensitivity to crack growth [16–18]. A change in the stress ratio will have a direct influ-
ence on crack closure, since a change in the stress ratio with a fixed stress range means that the mean stress is changed and
with it the plastic deformation at the crack tip. The effect of plasticity is, however, not directly considered in linear-elastic

Fig. 4. A part of the global model and the location of the defined stress.
M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163 155

fracture mechanics; instead, this effect is included by using semi-empirical formulas considering the elastic stress state at
the crack tip during crack propagation. The reduction in the crack growth rate due to the effect of crack closure is accom-
modated by introducing an effective stress intensity range when applying linear-elastic fracture mechanics, as shown in
Eq. (1)
DK cc
eff ¼ K max  K op ð1Þ

In this study, however, the effect of crack closure caused by cyclic loading has been accounted for only ‘‘globally’’ by
excluding the compressive stresses (the crack remains closed) from the variable amplitude fatigue loading. The effective
stress intensity factor was defined in Eq. (2) as follows:
DK cc
eff ¼ K max  maxðK min ; 0Þ ð2Þ

In the variable amplitude fatigue loading analyses (the VA analyses), the crack growth rates for any given crack front point
are computed using the following formulas [19]:
DK i ¼ DK max;i  DK min;i ð3Þ
when considering the crack closure globally:
DK min;i ¼ maxðDK min;i ; 0Þ ð4Þ

Ri ¼ K min;i =K max;i ð5Þ

da
!
dNi
ðDK i ; Ri ; . . .Þ
Dai ¼ Damedian da
ð6Þ
dNi
ðDK median ; Rmedian ; . . .Þ

where i is the load range, n is the number of load ranges in the spectrum and Ntot is the total number of load ranges. The
estimated fatigue life based on the average crack growth rate and random load history is a function of crack size, which will
be recomputed for each crack increment.

3.3. Fracture mode conditions

In fatigue-loaded structural details, the loading mode may vary from pure mode I, mode II and mode III to mixed-mode
conditions. Due to the complexity of loading and geometry in the studied detail, the fatigue-damaged region is subjected to a
mixed-mode fracture. It is reported in [20] that distortion-induced fatigue cracking is ‘‘generally’’ a result of mixed-mode
loading effects; mode I (opening) and mode III (tearing). In the fatigue crack-propagation analyses performed in the current
work, two alternatives have been considered: (1) a dominant mode I condition, i.e. DKI as the governing crack-opening mode,
and (2) mixed-mode loading conditions using DKeff, considering modes I, II and III. The effective stress intensity factor is
computed by using a combination of all three modes. The DKeff is calculated following the recommendation given by
BS7608:1999 [21] based on the strain energy release rate in Eq. (7).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
DK 2III
DK eff ¼ DK 2I þ DK 2II þ ðm ¼ 0; 3Þ ð7Þ
1m

3.4. Crack-propagation direction – kink angle

The crack growth direction has a great impact on the computed stress intensity factors and is directly dependent on the
stress conditions at the crack front. In the event of mode-I conditions, the crack will grow in its original crack plane, i.e. pla-
nar crack propagation. However, in complex cases such as the studied detail on the Söderström Bridge, where the crack is
under a complex stress condition, the crack growth direction is not obvious and not easy to predict. In order to compare
the crack size and direction in the analysis with the real cracks detected in the Söderström Bridge, three common crack
growth direction criteria were applied to determine the kink angle of the crack path. These criteria are the maximum tensile
stress criterion (MTS), the maximum generalised stress criterion (GEN), and the maximum strain energy release rate crite-
rion (SERR). All three criteria were available in Franc3D v6 [19].

3.5. Estimation of fatigue life

The fatigue life of the studied detail based on the computed stress intensity factors obtained from the performed crack
growth analyses was estimated by using a crack growth relationship proposed by Paris and Erdogan, known as the Paris
law [22], given in Eq. (8):
da
¼ C  DK m ð8Þ
dN
156 M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163

The Paris law parameters C and m were defined by following the recommendations given in BS7608:1993 [21], with
C = 5.21  1013 N/mm3/2 and m = 3. The threshold stress intensity factor (DKth) used for assessing welded joints was
63 N/mm3/2 recommended by BS7910:1999 [21] and IIW [23] was applied in this study.
In the crack-propagation analyses, the stress intensity factors were computed by defining the average crack growth rates
for the variable amplitude fatigue load. The total service life or residual service life in the event of damaged structures can be
calculated by following the linear damage model which is the most common method for adding up the fatigue damage, as
shown in Eqs. (9)–(12).

da 1 X n
da
¼ ðDK i ; Ri ; . . .Þ ð9Þ
dN Ntot i¼1 dN i

     
da da da
af ¼ a0 þ  N1 þ  N2 þ    þ  Ni ð10Þ
dN 1 dN 2 dN i

where af is the final crack size and a0 is the initial crack size.
The fatigue life can be computed by integrating the average crack growth rates as shown in Eq. (11).
Z af  
dN
N¼ da ð11Þ
a0 
da
For mixed-mode conditions when using the effective stress intensity factor range considering modes I, II and III, a
modified Paris law shown in Eq. (12) was used. The same equation has also been used for the effective stress intensity factor
range when the crack closure mechanism was considered. The only difference here is that the effective stress intensity
factors were calculated with and without taking the crack closure effects into consideration.
da  m
¼ C  DK eff ð12Þ
dN
The various types of crack-propagation analysis depending on the crack growth model and crack direction criteria performed
in this study are summarised in Table 1.

4. Results and discussion

4.1. Results of the VAFL analysis

The results of the crack growth analysis with the three different crack direction criteria, while taking the effect of com-
pressive stresses into consideration, revealed that the crack growth rate accelerates until the crack reaches a certain length.
At this point, when the deformation, i.e. out-of-plane distortion causing the crack initiation and propagation, is relaxed, the
crack growth rate subsequently decelerates. This was the dominant behaviour obtained from all the crack-propagation
analyses and it was also expected as a result of the propagating crack in the welded detail.
The crack growth path obtained from the variable amplitude load with the MTS criterion in the studied cracked region is
shown in Fig. 5. The result is highly consistent with the real crack formation. As seen in this figure, the crack growth is not
symmetrical on both sides of the stiffener. As mentioned earlier, the crack was initiated and propagated by out-of-plane dis-
tortion at the web plate caused by the end rotation of the cross-beam. Since the beam was positioned at an 80-degree angle
with reference to the main girders (on the left in Fig. 5), the crack formation on the left (narrow angle side) is diagonal, while
the crack formation on the right is more horizontal in the web plane. For the same reason, the crack growth in the thickness
direction (inside the web plate of the main girder) is non-planar and propagates downwards, as shown in Fig. 6b.
The computed stress intensity factors in the thickness direction (SIFt) and in the web plane (SIFLeft and SIFRight) from the
crack propagation analysis with MTS are shown in Fig. 6a. The crack formation and growth rates are different on the left- and
right-hand sides of the cross-beam along the plane of the web of the main girder, for the above-mentioned reason, while the
crack growth behaviour is identical. The only difference in the crack growth behaviour between the left- and right-hand sides

Table 1
Investigated various types of crack propagation analysis.

Crack growth model Kink angle criterion Type of crack propagation Crack growth governing mode
VAFL without crack closure MTS Specified Da KI mode I
Keff (mixed-mode)
GEN Specified Da KI mode I
Keff (mixed-mode)
SERR Specified Da KI mode I
Keff (mixed-mode)
VAFL with crack closure MTS Specified Da KI mode I
Keff (mixed-mode)
M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163 157

Fig. 5. Simulated crack growth from the VAFL analysis and a detected crack in the bridge.

Fig. 6. The stress intensity factors in the thickness (SIFt) and longitudinal directions on the left- (SIFLeft) and right-hand (SIFRight) sides of the crack and crack
shape inside the web plate.

is that the crack growth rate on the left-hand side is slightly faster. According to the values of stress intensity factors shown
in Fig. 6a, it is obvious that the crack growth rate in the thickness direction decreases rapidly and the crack will probably be
arrested in this direction. Moreover, Fig. 6 also shows that the crack growth in the web plane direction will continue at a
lower rate, even though the crack growth rate decreases rapidly over the weld toe line.
Comparing the results of the analyses with mode-I and mixed-mode loading conditions, it can be seen that the difference
is insignificant. This result basically reveals that mode I has an unquestionably dominant effect on crack propagation, even
though this propagation is the result of a combination of modes I, II and III. This observation is consistent with the results
reported in [1]. The domination of mode-I crack growth is clearly evident for crack propagation both in the thickness direc-
tion and in the direction of the crack path in the plane of the web.
The results for the different crack growth criteria considered in this study showed very similar crack development, see
Fig. 7. As stated in [1], the governing mode, especially for the crack along the weld toe line over the stiffener, is the opening
mode and the effect of shear and twisting is negligible in this region. Due to this insignificant shear effect, the different crack
direction criteria yield the same results. Since the shear and twisting effect is more visible when the crack grows outside the
stiffener region, the difference from the different crack direction criteria is more noticeable, as shown in Fig. 7.
Despite the fact that different models show slightly different crack-propagation paths, the crack-propagation rate is very
similar with the GEN and SERR, see Fig. 8a and Fig. 8b. These analyses confirm the accuracy of the results of the analyses
presented above.

4.2. Results of the crack closure analysis

The result of the crack-propagation analyses, when ignoring the negative portion of the load cycle resulting in crack sur-
face contact, is shown in Fig. 9. The computed stress intensity factors both for mode-I and for mixed-mode conditions are
almost identical, as shown in Fig. 9. This was expected, as the crack can only grow for the part of the tensile stresses in
the load history (shown in Fig. 2), i.e. when crack surfaces are open. The results are very similar, even though there are very
few differences between the SIF values for the extended crack in the web plane direction. This may be explained by the
meshing or template radiuses which were not exactly the same for all the analysis steps. On the other hand, the effects
of crack closure on crack growth could be captured by only considering the plasticity at the crack tip. In this case, the crack
tip is closed even when the applied stresses are slightly positive.
The influence of residual welding stress was not considered in the crack-propagation analysis due to the lack of knowl-
edge of the residual stress field in the studied detail. The residual stress may have a large influence on the fatigue life, which
158 M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163

Fig. 7. Comparison of various crack direction criteria when considering KI.

Fig. 8. The stress intensity factors in the thickness (SIFt) and longitudinal directions on the left- (SIFLeft) and right-hand (SIFRight) sides of the simulated crack
for the analysis with GEN and SERR criterions.

will be reduced in the event of residual tensile stress and increased in the presence of residual compressive stress. On the
other hand, the residual welding stress in the studied detail would not influence the crack growth rate completely, as the
crack propagated away from the welded region both in the thickness direction and along the web plane. This means that
the effects of the residual welding stress on the crack growth rate would only be limited to the welded region.
The ‘‘fatigue life’’ of the studied detail for all the studied cases is presented in Table 2, where it was computed using the
crack-propagation analyses for a certain length presented in Table 2. The crack length in the thickness direction was chosen
as 17 mm, while the crack length in the longitudinal direction was chosen as 30 mm for the left-hand side and 36 mm for the
right-hand side. The reason for using these crack lengths is to be able to compare the results from variable amplitude anal-
yses with those from constant amplitude analyses presented in [1]. According to these results, it would take almost 20 years

Fig. 9. The stress intensity factors in the thickness (SIFt) and longitudinal directions on the left- (SIFLeft) and right-hand (SIFRight) sides of the crack and crack
shape inside the web plate.
M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163 159

Table 2
Total fatigue life of the studied detail according to various crack propagation analyses.

Crack growth model Kink angle criterion Crack propagation Crack mode Fatigue life for crack length of
x = 17 mm
z = 30 mm (right)
z = 36 mm (left)
VAFL without crack closure MTS Specified Da KI 11,023,962 (20.5 years)
8,781,892 (16.4 years)
9,261,176 (17.3 years)
Keff 10,722,716 (20.0 years)
8,937,850 (16.7 years)
9,348,402 (17.4 years)
GEN Specified Da KI 10,623,601 (19.8 years)
9,297,537 (17.3 years)
9,775,967 (18.2 years)
Keff 10,198,798 (19.0 years)
9,299,642 (17.3 years)
9,448,679 (17.6 years)
SERR Specified Da KI 10,569,109 (19.7 years)
9,163,608 (17.1 years)
9,546,278 (17.8 years)
Keff 10,314,246 (19.2 years)
8,966,257 (16.7 years)
9,567,635 (17.8 years)
VAFL with crack closure MTS Specified Da KI 10,694,229 (19.9 years)
9,068,206 (16.9 years)
9,465,954 (17.6 years)
Keff 9,943,977 (18.5 years)
9,184,113 (17.1 years)
9,471,908 (17.7 years)

for this crack to develop for the given crack length. However, the first crack in the bridge was detected 50 years after the
bridge was built. This can be explained by the actual loads applied in the analyses which are higher than the ‘‘older’’ loads
and the fact that the traffic volume has increased substantially during this period.
The result of the crack-propagation analyses shows that the crack growth in the studied section will be arrested in the
thickness direction, as shown in Fig. 10a, while the crack growth in the web plane direction will continue to propagate,
but at a reduced crack growth rate, as shown in Fig. 10b. The turning point for crack growth, as seen in Fig. 10a, was when
the crack had propagated 7 mm in the thickness direction (see Fig. 6).

4.3. Comparison with constant amplitude fatigue loading

When comparing the results obtained in the crack growth analysis with variable amplitude loading (VA analysis) with the
results from the crack growth analysis with constant amplitude loading (CA analysis) presented in [1], it appears that the
crack propagation simulated in the VA analysis is faster than the crack propagation in the CA analysis, as shown in Figs. 11
and 13. In other words, the estimated fatigue life of the studied detail based on the VA analysis is shorter than the estimated
fatigue life from the CA analysis. It was expected that the VA analysis would yield a lower crack growth rate due to the inter-
action effects, i.e. the beneficial compressive stresses in the load history. If the comparison is made at the same equivalent
stress range level – the equivalent stress range calculated from the load history used in the VA analysis – the difference is
significant, as shown in Fig. 11. One reason for this difference between the crack growth rates might be the defined load in
the CA analysis. The studied welded detail is subjected to bending stress due to the end rotation of the cross-beam and the
detail category of C56 for longitudinal attachments recommended in EC3 is for tensile stress, i.e. S–N curve used for calcu-
lating the equivalent stress range. However, there is no explanation or recommendation for such details with bending. Using
a higher S–N curve would probably have resulted in a higher stress, due to a change in the proportion of stress levels
included in the calculations and thereby the shorter fatigue life.
There is yet another possible reason for this slower crack growth rate from the CA analyses and it is related to the cal-
culation of the equivalent stress range from the load history. As stated in [1], the equivalent stress range used in the CA anal-
ysis was computed using the Rainflow stress accounting method from the load history. The equivalent stress was calculated
by considering slopes of 3 and 5 in relation to the stress levels, i.e. a constant amplitude fatigue limit at 5  106 cycles and a
variable amplitude fatigue limit at the knee point at 1  108 cycles given in EC3 [3], see Fig. 12. The equivalent stress was
computed as 19.0 N/mm2, which was under the cut-off limit for a fatigue detail class of C56, which in turn gives an infinite
fatigue life. The limit for the constant amplitude stress within the S–N curve defined in EC3 influences the computed
equivalent stress substantially, see Fig. 12. For instance, in the IIW’s recommendations for the same fatigue strength curve
(56 N/mm2), the computed equivalent stress range would be computed as 26.1 N/mm2 and the results would then be a bet-
ter match for the results obtained from the VA analysis.
160 M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163

Fig. 10. Distortion-induced fatigue crack behaviour in the studied welded detail.

Fig. 11. Comparison of the results from the CAFL with equivalent stress and VAFL analysis.

When comparing the results obtained from the VA analysis with the results obtained from the CA analysis based on the
load from the X60 train type, the difference between the crack growth rates is not as large as that in the CA analysis with
equivalent stress, as shown in Fig. 13. This is again because of the defined bridge load in the analysis, as the stress range
caused by the X60 is higher than the equivalent stress range. Using the maximum load (apart from freight trains which pro-
duced few stress cycles) appears to be conservative. However, uncertainties in the equivalent stress calculations, such as the
fatigue constant limit for the S–N curve, the low stress levels that are not included in the calculations or the small number of
load cycles with very high stress levels produced by freight trains, can be covered using the maximum possible load range.
Moreover, the proportion of the load history from the X60 is fairly high and this can be confirmed by a Palmgren–Miner sum-
mation. When considering a total fatigue life, i.e. the Palmgren–Miner sum of D = 1, the stress range intervals produced by
the X60 train are about 63% of the total service life.
The main mechanism for taking account of the interaction effects (stress interactions) causing crack growth retardation is
the plasticity-induced crack closure (Elber mechanism) [18], which is not included in this study. The higher crack growth
rate from the VA analysis might be also dependent on ignoring this interaction effect, which could result in a retardation
of crack growth. However, the effect of load interactions was considered in the analysis, which would normally yield a longer
fatigue life with lower R ratios, which is the case in the VA analysis. The defined stress ratio was zero (0) in the CA analysis,
while the average R ratio when considering the load history in the VA analysis was 0.5. On the other hand, the effect of
mean stress is known to be insignificant for the fatigue strength of welded joints, due to the very high notch effect caused
by attached plates and welds. Kassner and Krebs [24] experimentally investigated the relationship between the notch effect
and the mean stress effect on the fatigue strength of welded joints. Their results showed that the notch effect substantially
reduces the effect of mean stress on fatigue life.
One uncertain parameter is the threshold value (DKth) used for the service life calculations. The threshold value was
assumed to be valid for both the CA and the VA studies. However, it is known that the threshold values are obtained by
means of experiments with a decreasing crack growth rate (not using random loading shown in Fig. 2) to define the crack
growth resistance which may not be representative of VA loading [18]. This implies that the crack-driving forces that are
smaller than the crack growth resistance can still contribute to crack growth.
M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163 161

Fig. 12. Fatigue strength curve of 56 according to EC3 (C56) and IIW (FAT56).

Fig. 13. Comparison of the results from the CAFL (loads from X60) and VAFL analysis.

4.4. Inspection interval

The replacement of existing bridge structures is generally expensive and can cause huge logistical problems on the rail-
way network. Repairing damaged components in bridges might be sufficient, depending on the condition of the bridge, and
thereby more economical. It is therefore important to understand the condition of damaged bridges, which may result in a
more effective repair schedule and fewer traffic disruptions.

Fig. 14. Recommended inspection strategy for the Söderström Bridge.


162 M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163

According to the results obtained in the study presented in this paper, an inspection strategy can be recommended for the
unrepaired distortion-induced fatigue cracks detected in the Söderström Bridge, as shown in Fig. 14. In this recommendation,
the first inspection interval corresponds to a crack length of 9 mm, while, for the second inspection interval, a shorter crack
length of 7 mm is expected, according to the analyses performed in this study. In the event of increasing traffic volumes in
the future, shorter inspection periods should be considered. This can easily be confirmed by detecting longer crack lengths in
the second inspection.

5. Conclusions

Distortion-induced fatigue cracking, which is a common source of fatigue damage due to unforeseen local component
behaviour, unintended or otherwise overlooked interaction between the bridge components, has been investigated by LEFM
with ‘‘real’’ bridge loads. Based on the analysis results and discussions presented in the previous section, the following con-
clusions can be drawn.

 The crack growth rate decreases with the extended crack length due to relaxation, as the web gap becomes more flexible
because of the presence of the propagated crack. The out-of-plane distortion causing fatigue cracking is also an indicator
of the retardation effect.
 The results confirm that the stress intensity factors have three different crack growth rates in this studied welded region.
 The crack direction criteria studied in this investigation provide basically the same crack formation and crack growth rate.
 The results of this study show that the effect of mean stress is less significant in the studied detail due to the presence of a
notch (crack tip).
 There is generally a significant difference between constant and variable amplitude fatigue crack growth analyses. The
main reason for this difference is the bridge load defined in the analyses. Variable amplitude fatigue crack growth analysis
produces more conservative results.
 The results underline the great importance of accurately defining the design load. Considering the worst case may provide
more a reliable fatigue life estimate and a not very high conservative life estimate, as the uncertainties in the calculations
and assumptions can be accommodated.

Acknowledgements

This investigation is being carried out within the research project funded by the Swedish Transport Administration
(Trafikverket) and this funding is gratefully acknowledged. The authors would also like to express their appreciation to Bruce
J. Carter and Paul A. Wawrzynek for their valuable support in performing the analyses in this study.

References

[1] Aygül M, Al-Emrani M, Borsoum Z, Leander J. Investigation of distortion-induced fatigue cracked welded details using 3D crack propagation analysis.
Int J Fatigue 2013;64:54–66.
[2] BS 5900: Part 10. Steel, concrete and composite bridges, Part 10. Code of practice for fatigue; 1980.
[3] EN 1993-1-9. Eurocode 3: design of steel structures – Part 1-9: Fatigue, CEN, Brussels; 2005.
[4] AASHTO LRFD Bridge Design Specifications (American Association of State Highway and Transportation Officials). 5th ed. Washington, DC; 2010.
[5] Kühn B, Lukic M, Nussbaumer A, Günther H-P, Helmerich R, Herion S, et al. In: Sedlacek G, Bijlaard F, Géradin M, Pinto A, Dimova S, editors. Assessment
of existing steel structures: recommendations for estimation of remaining fatigue life; 2008.
[6] EN 1991-2. Eurocode 1: actions on structures – Part 2: traffic loads on bridges. Brussels: European Committee for Standardization; 2003.
[7] Nussbaumer A, Borges L, Davaine L. Fatigue design of steel and composite structures. ECCS – European Convention for Constructional Steelwork, ECCS;
2011.
[8] Al-Emrani M. Fatigue-critical details in steel bridges. Department of Structural Engineering, Chalmers University of Technology, Report 2006:7; 2006
[in Swedish].
[9] Al-Emrani M. Fatigue in riveted railway bridges – a study of the fatigue performance of riveted stringers and stringer-to-floor-beam
connections. Gothenburg: Chalmers University of Technology; 2002.
[10] Al-Emrani M. Fatigue performance of stringer-to-floor-beam connections in riveted railway bridges. J Bridge Eng 2005;10:179–85.
[11] Al-Emrani M, Åkesson B, Kliger R. Overlooked secondary effects in open-deck truss bridges. Struct Eng Int J Int Assoc Bridge Struct Eng (IABSE)
2004;14:307–12.
[12] Fisher JW. Fatigue cracking in bridges from out-of-plane displacements. Can J Civ Eng 1978;5:542–56.
[13] Fisher JW, Fisher TA, Kostem CN. Displacement induced fatigue cracks. Eng Struct 1979;1:252–7.
[14] Leander J, Andersson A, Karoumi R. Monitoring and enhanced fatigue evaluation of a steel railway bridge. Eng Struct 2010;32:854–63.
[15] Elber W. The significance of fatigue crack closure. ASTM Spec Tech Publ 1971:230–42.
[16] Andersson TL. Fracture mechanics: fundamentals and applications. 3rd ed. New York, USA: CRC Press Taylor and Francis Group; 2005.
[17] Zhang S, Marissen R, Schulte K, Trautmann KK, Nowack H, Schijve J. Crack propagation studies on AI 7475 on the basis of constant amplitude and
selective variable amplitude loading histories. Fatigue Fract Eng Mater Struct 1987;10:315–32.
[18] Schijve J. Fatigue of structures and materials. Netherlands: Springer; 2009.
[19] Franc3D. Reference manual for version 6. Fracture Analysis Consultants Inc; 2011.
[20] Fraser REK, Grondin GY, Kulak GL. Behaviour of distortion-induced fatigue cracks in bridge girders. Structural engineering report 235. Department of
Civil and Environmental Engineering, University of Alberta, Alberta; 2000.
[21] BS7910:1999. Guide on methods for assessing the acceptability of flaws in fusion welded structures. London, United Kingdom: British Standard
Institution; 2000.
M. Aygül et al. / Engineering Failure Analysis 45 (2014) 151–163 163

[22] Paris P, Erdogan F. A critical analysis of crack propagation laws. J Basic Eng 1963;85:528–33.
[23] Hobbacher A. Recommendation for fatigue design of welded joints and components. Paris, France: The International Institute of Welding; 2008.
[24] Kassner M, Krebs J. Influence of welding residual stresses and notch effect on fatigue data for welded joints and components. Germany: International
Institute of Welding; 2007.

Potrebbero piacerti anche