You are on page 1of 10

Journal of Constructional Steel Research 119 (2016) 113–122

Contents lists available at ScienceDirect

Journal of Constructional Steel Research

Experiment on fatigue behavior of rib-to-deck weld root in orthotropic


steel decks
Shigenobu Kainuma a, Muye Yang a, Young-Soo Jeong a,⁎, Susumu Inokuchi b,
Atsunori Kawabata b, Daisuke Uchida b
a
Department of Civil Engineering, Kyushu University, Fukuoka 8190395, Japan
b
Technical Research Laboratory, Japan Bridge Association, Tokyo 1050003, Japan

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

Article history: This study experimentally investigated the fatigue behavior of the weld root in orthotropic steel decks stiffened
Received 21 August 2015 with U-ribs in relation to the loading conditions and welding details. To examine the structural response and local
Received in revised form 24 November 2015 stress near the welded joint, field loading tests, measurement of residual stress, and fatigue tests were carried out.
Accepted 26 November 2015
A total of 12 specimens were manufactured, and the fatigue tests were performed by simulating the double tire
Available online 17 December 2015
loading of an actual vehicle. The welding residual stress distribution at the root was examined to understand the
Keywords:
mechanism of root crack initiation and propagation behavior. Thus, fatigue behavior of the root crack was inves-
Orthotropic bridge deck tigated and evaluated by considering different stress ratios and weld penetration rates. Based on the fatigue test
Root crack results and crack patterns, it was revealed that both tensile stress and stress range could affect the root crack ini-
Fatigue test tiation. However, the tensile stress rather than the magnitude of stress range would be the effective stress after a
Residual stress crack initiated, and it was seen to be an important factor for root crack propagation. In addition, a penetration rate
Weld penetration in the range of 0% to 75% was beneficial for fatigue durability of this structural detail.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction were many other types of cracks. However, crack initiation at the weld
root of the longitudinal fillet welding between deck plates and closed-
Orthotropic steel bridge decks were widely applied to long-span rib cannot be observed by visual inspection, which might have serious
bridges because of their characteristics such as light weight, high implications to bridge safety. Moreover, at present, U-ribs account for
strength and durability, and rapid construction [1,2]. In Japan, about 60% of rib stiffeners in the orthotropic steel deck construction.
many orthotropic steel bridges were extensively constructed in the The rib-to-deck structure includes a large number of single-fillet weld
1960s–80s. Recently, various types of damage damages to the steel joints because of the wide use of U-ribs [5]. Therefore, fatigue cracks
decks were reported owing to long-term operation and the changed op- of the weld root between the deck plate and U-rib are one of the most
erating environment including the emergence of heavy-duty vehicles common and serious cracks, because of the inadequate welding details
and increasing traffic volume. in conjunction with the residual stress induced by the welding process
Fatigue cracks in orthotropic steel bridges significantly occurred at and unfavorable weld shape.
partial-penetration fillet-welded connections owing to the cyclic load The root cracks have been investigated by several experimental and
stress caused by a high number of vehicles. The following typical fatigue analytical studies. Analytical studies usually focused on the local stress
cracks have been reported in recent years [3], as shown in Fig. 1. A rib- around the rib-to-deck weld joint, and the fatigue durability of the
to-deck crack is usually initiated at the weld root, and it propagates structure was evaluated by finite-element simulation [6]. Past experi-
into the deck plate or through the weld bead. It has been reported mental studies were conducted through full-scale fatigue tests under
that root cracks in U-rib to deck connections account for about 18.9% fixed-point loading [7–9] and rib-to-deck small specimen tests [10,11].
of the total damage types in orthotropic steel decks [4]. The weld toe Full-scale fatigue tests of orthotropic steel decks are usually conducted
crack of the deck plate, which is located at the rib-to-cross beam con- to evaluate the fatigue performance. Bending fatigue tests of small spec-
nection with a partially welded joint, and the crack occurring at the fillet imens can be conducted to evaluate the fatigue life of crack initiation
weld toe near field-welded splice plates fixed to the webs were also and propagation within a certain length of the specimens, and it is
common damage types in orthotropic steel bridges. In addition, there possible to assess the fatigue crack growth and the fatigue strength.
However, the boundary condition is different between the actual
orthotropic steel decks and small specimens. It is difficult to compare
the fatigue behavior of small specimens with that of actual orthotropic
⁎ Corresponding author. steel decks. This is because the tests are performed under constant

http://dx.doi.org/10.1016/j.jcsr.2015.11.014
0143-974X/© 2016 Elsevier Ltd. All rights reserved.
114 S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122

Table 1
Dimensions of the orthotropic steel decks of A1 Bridge (unit: mm).

Thickness of asphalt pavement 75


Thickness of deck plate 12
Dimension of U-rib 340 × 284 × 8
Spacing of the transverse rib 3000
Spacing of U-rib 680
Spacing of web of girder 2800
Distance between diaphragms 5600

Based on A1 Bridge, this study focused on the rib-to-deck weld joint


where the root cracks had been observed. A field loading test was con-
ducted in August (2005) when the influence of the asphalt pavement
was most negligible at high temperatures. The mid-span of the U-rib
was set as the target section, and the measured transverse stresses
near the target used as “reference stress”, which were 5 mm away
from the weld toe at the bottom of the deck plate. The measured refer-
ence stress range was about − 40 MPa to 6 MPa under the rear tires,
as shown in Fig. 2. Although the tensile stress was much lower than
the compressive stress, it has the potential to open the weld root gap,
Fig. 1. Typical fatigue cracks in orthotropic steel bridges.
and lead to the initiation and propagation of the root cracks [9]. In
this study, a stress ratio σt,max/σc,max equal to 0.13 was obtained
from field measurement data. The stress range of A1 Bridge was deter-
mined by a stress histogram measurement, which was carried out for
strain range and without restraint on the rib. Moreover, the secondary 3 days by using the rain-flow method, and the maximum stress range
stress caused by deformation and the effect of release of the welding was 170 MPa [14,15]. These results were used in the following fatigue
residual stress after crack initiation could not be properly simulated in tests.
these tests.
For penetration rate, most researchers considered that a higher pen- 3. Fatigue test
etration rate could lead to a higher strength in the weld joint. However,
the welding process of full penetration weld is more complex. Its reheat 3.1. Test specimen
cycle would reduce the performance of the welded joint and easily
cause weld defects. Recently, full-scale fatigue tests of orthotropic A total of 12 fatigue tests were conducted on full-scale orthotropic
steel decks conducted by some researchers have proved that a steel deck specimens. Each specimen was 1400 mm wide and consisted
shallower weld penetration at the partial joint penetration (PJP) joint of two U-ribs of 2000 mm span between two cross beams. Steel grade of
appeared to have a positive effect on enhancing the fatigue resistance. JIS G3106 SM400A [16] was used for the tested specimens, whose
Rib-to-deck fatigue tests of small specimens also showed that the spec- chemical composition and mechanical properties are listed in
imens with weld melt-through seemed to have slightly lower fatigue Table 2(a). Each specimen consisted of a 12-mm-thick deck plate and
strengths than those with 80% PJP [11,12]. Although most researchers 6-mm-thick U-rib, and was welded by a semi-automatic CO2 welding
agree that a penetration rate below 75 ~ 80% is beneficial to fatigue du- method. Three types of specimens were fabricated with 0%, 75%, and
rability, it is still inconclusive whether full penetration could enhance 100% penetration rates at the U-rib to deck plate connection. The di-
the fatigue strength of a rib-to-deck weld joint. mensions and welding conditions of these specimens are listed in
In this study, the fatigue behavior at different penetration rates at the Table 3, where D and U are the thickness of the deck plate and U-rib, re-
PJP weld joint was experimentally evaluated. The structural responses of spectively, and SP is the penetration rate. Thus, D12U6SP0, for example,
the reference points near the weld joints were examined using field mea- represents a 12-mm-thick deck plate and 6-mm-thick U-rib with a weld
surements and fatigue tests. The residual stress at the weld root of the joint of 0% penetration rate. Photographs of the weld joints of the spec-
deck plate to U-rib connection was measured using a cutting method. imens in the etching test are shown in Fig. 3; in the figure, the heat af-
The test results also explained the cause identification of the crack initia- fected zone caused by the welding process can be observed. The
tion of this structural detail. In addition, based on the measurement of the welding area varies between the specimens with different penetration
welding residual stress and fatigue test results, the fatigue behavior and rates.
cracking mechanism of the weld root between the deck plate and U-rib The dimensions of the specimens and the loading positions are shown
were investigated by considering the effective stress. in Fig. 4. The dynamic loading area at the mid-span was based on the ac-
tual double tires. As the weld root tip stress could not be measured during
2. Loading test of orthotropic deck bridge the test, the transverse stress of points that were 5 mm away from the
weld toe in this weld line was considered as the reference stress. The
A1 Bridge is located in the heavy traffic route of the Tokyo Metro bay stress at the weld root tips and the reference points tend to have a positive
area in Japan. A1 Bridge consists of two simple box girder bridges and a correlation. The tested primary reference stress at the mid-span was con-
three-span continuous box girder bridge with orthotropic steel decks. sidered as the stress condition for each specimen.
Large vehicles account for 45.2% of the daily traffic volume. The dimen-
sions of the orthotropic steel deck are given in Table 1. Asphalt pave- 3.2. Test set-up
ment damage in A1 Bridge was visually inspected in 2004, and severe
root cracks were observed during the re-paving construction process The experimental set-up consisted of two hydraulic jacks that con-
in 2005. The root cracks occurred around the intersections of the U-rib trolled static loading Ps and dynamic loading Pd. The loading positions
at the weld joints between diaphragms, and the maximum length of are shown in Fig. 5. Ps could be changed to vary the stress ratio so that
cracks was over 400 mm [13]. both the tensile stress and compressive stress could be simulated. Pd
S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122 115

Fig. 2. Field measurements in loading test of orthotropic deck bridge.

provided a cyclic load, and its hydraulic jack applied force to the loading and tire area of the single test were set at half the corresponding values
jig. The loading jig consisted of a loading brick, steel plate, and rubber of the double tire test. A single loading plate was placed between two U-
plate, as shown in Fig. 6. Two steel plates of 25 mm thickness were ribs at the mid-span.
used as the double tires, whose size depended on the pressure area of
the actual double tires [14]. The fatigue tests were conducted using rub-
ber plates instead of asphalt pavement. The properties of the rubber 3.3. Test results
plates are summarized in Table 2(b). This study used rubber plates of
80 mm thickness so that load dispersion could be considered effectively The conditions of test stress, crack length, and crack shapes are given
[15]. in Table 4. The surface cracks were detected by the magnetic particle
Based on field measurement results of A1 Bridge, the test primary testing (MPT) method after all the tests were completed. The cracks
stress range of the mid-span reference point of basic specimens was mainly existed at a distance of ± 200 mm from the mid-span. It was
controlled at 180 MPa. The basic fatigue test set the maximum compres- thought that the cracks were mostly initiated at the mid-span and
sive stress at −160 MPa and the maximum tensile stress at 20 MPa be- extended in both depth and length directions. Therefore, the specimens
cause the measured stress ratio of A1 Bridge was 0.13. The stress ratio R were cut into 25-mm-thick pieces in the longitudinal direction, so that
is defined as the algebraic ratio of absolute value of minimum stress to the crack shape and details in the cut sections could be acquired.
maximum stress. The other stress range conditions and different stress The photos of the crack patterns are shown in Fig. 8. The root crack
ratios were used for comparison tests. The maximum value of tensile fracture surface of D12U6SP0-4 in the longitudinal direction is shown
stress was considered as an important factor for crack initiation, and in Fig. 8(a). The fracture surface consists of some discontinuous
the effects of stress ratio for a given stress range were explained, fractures that could be multi-cracks initiated by the root crack and
e.g., the stress range of −140 MPa to 40 MPa. eventually forming a large crack. The corresponding crack patterns
Test coordinates were set at the mid-span of longitudinal U-rib as of D12U6SP75 and D12U6SP100 by MPT inspection are shown in
x = 0, and the static load side was regarded as the positive x direction. Fig. 8(b, c, d).
Uniaxial strain gages were attached at the reference points at a distance The test results could be divided into three crack patterns as follows:
of 150 mm to 600 mm away from the mid-span. The test primary stress (1) Only root crack existed; (2) Both toe crack and root crack existed,
distribution in the longitudinal (x) direction under dynamic loading is and at N ar; (3) No crack initiated. The crack patterns are described in
shown in Fig. 7. Table 5.
Three million cycles were applied to all specimens in the first loading An analysis of the stress condition of category (1) revealed that the
stage. If there was no crack initiation, another 1.5 million cycles were tensile stress and crack length of the SP0 specimen were larger than
applied to continue the testing. An additional single tire loading was ap- those of the SP75 and SP100 specimens for the same stress range.
plied to two D12U6SP100 specimens, which was controlled by one hy- However, the stress range and crack depth of the SP75 specimens did
draulic jack that provided Pd at the mid-span. Both the dynamic loading not show a correlation because of the differences in the geometry of
the root gap. The root crack length of SP100 and SP0 were almost half
of deck thickness under the stress condition of − 140 to 40 MPa.
SP100-2 and SP100-3 were under the same stress conditions of
Table 2 first stage loading, and then, SP100-2 continued 3 million cycles of
Chemical composition and mechanical properties. single tire loading while SP100-3 continued 6 million cycles. Crack
depth comparison of two specimens showed their propagation lengths
(a) Specimen of fatigue test and measurement of residual stress
in the vertical direction were similar. Due to the limit of test number,
Yield point Tensile strength Elongation C Si Mn P S the effect of single tire load on the root crack propagation seems not
(MPa) (MPa) (%) (%) (%) (%) (%) (%) obvious so far. In addition, the case of SP100 showed a different
302 423 30 0.16 0.05 0.78 0.014 0.004 mechanical mechanism because of its special geometry of the root
gap at the weld joint. Stress concentration caused by the geometry
(b) Rubber plate mutation and the crack initiation point were always located at the
Yield point Tensile strength Elongation root tip of PJP, while stress distribution and stress concentration of the
(MPa) (MPa) (%)
full penetration weld joint were different because of the weld shape.
An alternative idea is that the SP100 specimen might have created a
1.2 3.2 250
geometrical discontinuity at the weld root that led to a lower fatigue
116 S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122

Table 3
Specimen dimensions and welding conditions.

Thickness (mm) Penetration Penetration rate Current Voltage Speed


Specimen Number of passes
Deck td U-rib tu Pw (mm) Pw/tu (%) (A) (V) (cm/min)

D12U6SP0 12 6 0 0 220–280 28–30 30 2


D12U6SP75 4.5 75 280 32 45 2
D12U6SP100 6 100 200–280 25–30 35–45 2

resistance. Related research can be found in the fatigue tests conducted deck, even for specimens with different penetration rates. In particular,
by Sim et al. [8]. residual tensile stress always existed at about 10% of the thickness of the
An analysis of the toe and root crack patterns showed that the crack deck plate from the root tip. Based on the residual stress test results
was easily initiated at the weld toe under a high stress ratio. Some past shown in Fig. 10, the values of residual stress near the root tip were con-
full-scale orthotropic steel deck fatigue tests on 12- and 14-mm-thick fined to 60 to 100 MPa, while the actual residual stress of the root tip
deck plates also showed both root and toe cracking. However, there should be higher. Superposition of the tensile stress and residual stress
has been no report of toe cracking in actual orthotropic steel decks would make the structure susceptible to fatigue failure [18]. However,
with 12 mm deck plates in Japan. The past full-scale orthotropic steel the residual tensile stress would be released after crack initiation and
deck fatigue tests proved that toe cracking would also occur with a propagation.
high tensile stress [9,17]. The crack depth of each specimen was usually smaller than half of
An analysis of category (3) showed that compressive stress would deck thickness, which agrees with the results of a previous study
not cause fatigue cracking even under the basic stress range of [7–9]. The full-scale tests with the 12-mm-thick deck plate showed
180 MPa. At the same time, a stress condition with half or 2/3 of the that the root-to-deck crack was likely to take a long time before coales-
basic stress range may lead to an improvement in fatigue life. This cence and failure, even though it nearly reached the surface. The full-
means that both stress ratio and stress range could affect the fatigue scale tests of the 16-mm-thick deck plate showed that the fatigue cracks
life before crack initiation. seemed to be concentrated at around half of the deck plate thickness.
Past tests with a 12- or 14-mm-thick deck plate also reported that the
4. Measurement of residual stress root-to-deck cracks mostly stopped propagating after they reached
the neutral axis of the deck plate thickness. As suggested before, all
The welding process creates high residual stress in the weld joint. these fatigue tests showed that the residual tensile stress would be re-
The magnitude of the peak residual stress may correspond to the yield leased after the root crack was initiated. The root crack appeared to
point. Thus, it would seem that the stress range, and not the magnitude close after the residual stress was released because of crack growth,
of the structural stress, determines the fatigue life. Even if the structural which finally stopped the crack propagation.
stress range is entirely compressive, the effect of the welding residual Therefore, the tensile stress instead of the stress range tends to affect
stress makes the structure susceptible to fatigue failure [18]. As shown crack propagation. However, the tensile stress was small for continued
in Fig. 9, the residual stress can increase the range of effective stress in- crack propagation. A past fatigue test on fillet-welded cruciform joint
tensity factor above the threshold value of K [19]. under compressive cyclic stress also concluded that the fatigue root
The residual stress distribution near the weld root tip could be mea- cracks propagated to a length of 4 mm (1/3 of plate thickness) and
sured by the cutting method [20,21]. To evaluate the welding residual then ceased to grow [22]. It was suggested that either the tensile stress
stress in the deck vertical direction, the residual stress of D12U6SP0-1 around the crack tip was not large enough to reach the stress intensity
was measured. The location of the cut out pieces of the specimen factor or the stress range provided only by the tensile stress was smaller
is shown in Fig. 4. In this study, to understand the distribution of resid- than the constant amplitude of the fatigue limit stress range, which
ual stress, residual stress tests using the cutting method were carried stopped the fatigue crack propagation [24].
out for three-scale large models of D36U18 with penetration rates of
0% and 75%. The residual stress was then compared with that of 5. Fatigue behavior of weld root
D12U6SP75.
The stress distribution of the tested three-scale specimens was sim- 5.1. Crack initiation and propagation process
ilar to the results from the full-scale fatigue test [20], as shown in Fig. 10.
Although residual stress values were different for each specimen, it was In order to investigate the fatigue behavior of root cracks, cracking
obvious that stress trends were similar in the vertical direction of the position, depth, and propagation angles were analyzed. Figs. 11 and 12

Fig. 3. Weld joint photos of specimens in etching test.


S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122 117

Fig. 4. Fatigue loading area and dimensions of specimen.

show the measured reference stress and relative reference stress range σ y −σ max;t;m ≤σ res Mid‐span ðMaximum stress range pointÞ ð3Þ
of the SP75 specimens. These figures show that the maximum stress
range and maximum compressive stress of all specimens were located where σres is the residual stress of the weld root tip, σy is the yield
in the mid-span, and the maximum tensile stress point should be be- strength of base metal, σmax, t and σmax, c are actual maximum tensile/
tween the dynamic and static loading positions. According to the resid- compressive stress at the root tip, respectively, and σmax,t,m is the
ual stress measurement and analysis, the theoretical crack initiation maximum tensile stress at the mid-span.
point should obey the following formulas: Crack images illustrated in Table 4 show that most of the root cracks
were initiated at the mid-span, which was consistent with the case of
σ res ≤σ max;c Maximum tensile point ð1Þ (3). The crack depth distribution of the D12U6SP0 specimens was dif-
ferent from others. The D12U6SP0-1 specimen had a crack depth in
σ max;c ≤σ res ≤σ y −σ max;t Maximum tensile point ð2Þ the range of 0 to 175 mm in the longitudinal direction. However, the
D12U6SP0-4 specimen had a crack depth in the range of − 200 to
200 mm, because the origin of fatigue failure changed with the magni-
tude of the stress range owing to the welding residual stresses in the
weld root [23].
The residual stress of the D12U6SP0-1 specimen seemed to be
smaller than those of the SP75 and SP100 specimens. Thus, the

Fig. 5. Test set-up.

Fig. 6. Loading jig. Fig. 7. Reference stress range of D12U6SP0.


118 S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122

Table 4
Test stress and results.

Stress (MPa)
Specimen ID Cycle (×104) Maximum crack depth, a (mm) Image of crack depth (mm) and shape
σmax σmin Δσ

D12U6SP0 1 20 −160 180 300 ar = 3.2

2 20 −100 120 450 No crack

3 20 −70 90 300 No crack

4 40 −140 180 300 ar = 5.4

D12U6SP75 1 20 −160 180 300 ar = 1.5

2 20 −100 120 300 ar = 2.7

3 40 −140 180 300 ar = 3.2


at = 4.5

4 0 −180 180 450 No crack

5 40 −140 180 300 ar = 4.6


at = 5.7

D12U6SP100 1 40 −140 180 300 ar = 2.5


at = 3.5

2 40 −140 180 300 ar = 6


+300Single

3 40 −140 180 300 ar = 6.1


+600Single
S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122 119

Fig. 8. Typical crack patterns of test results.

crack was initiated at the maximum tensile point and not at the In the case of SP0-4, the root crack had a relatively small depth at
mid-span. In contrast, the long cracks of D12U6SP0-1 were ob- the location where the deck crack propagated rapidly. The crack
served in the range x = 50 to 150 mm. To explain the effect of stress occurred on the upper surface of the deck plate because the steel
distribution according to the loading condition, the crack depth was reached the yield strength during the fatigue test. This crack was per-
measured in the y and z directions in each cut section as shown in pendicular to the deck plate, and a semi-elliptical crack was observed
Fig. 13. The red and blue lines indicate that the trends of root as shown in Fig. 15. It was thought that this type of deck crack tends
crack propagation were opposite, and the crack depth in the z direc- to propagates through the root crack. From the perspective of bridge
tion varied notably even for a given y value. The width of root cracks structural safety, this case would be worse than when only root cracking
in the transverse direction for different propagation angles is illus- occurs.
trated in Fig. 14. It is clear from Fig. 7 that crack angles at x = 125
were very different from those at other sections because the tensile 5.2. Crack behavior and parametric analysis
stress was larger than those for the other sections. The fatigue
cracks were usually observed as small semi-elliptical cracks at mul- In order to investigate the effect of penetration rate, crack behaviors
tiple points along the weld line, and then the small fatigue cracks of specimens SP0-5, SP75-3, SP75-5, and SP100-3 were compared.
propagated and coalesced into a larger crack. Thus, the x = 125 sec- These four specimens were under the same stress range (− 140 to
tion was probably near the location of the confluence of two small 40 MPa) and same fatigue life of 3 million cycles. Typical macro sections
cracks. of the root crack of 80% PJP and weld melt-through specimens in

Table 5
Relationship between stress situation and crack length.

Crack types Category No. Tensile stress Stress range Crack length
(MPa) (MPa)
ar (mm) at (mm)

Root crack only SP0-1 20 180 3.2 –


SP0-4 40 180 5.4
SP75-1 20 180 1.5
SP75-2 20 120 2.7
SP100-2 40 180 6
(+300Single)
SP100-3 40 180 6.1
(+600Single)

Toe and root SP75-3 40 180 3.2 4.5


cracks SP75-5 40 180 4.6 5.7
SP100-1 40 180 2.5 3.5

No crack SP0-2 20 120 – –


SP0-3 20 90
SP75-4 0 180
120 S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122

Fig. 9. Effective stress at the weld joint.

reference [24] showed that the fatigue crack paths are nearly perpendic-
ular to the deck plate. However, in this study, the root cracks were not
Fig. 11. Measured reference stress of D12U6SP75.
entirely perpendicular to the deck plate, which might have been due
to the different deck plate thickness and boundary condition. The root
cracks of all these tested specimens initiated at the root tip. The propa-
gation direction tends to approach the vertical direction of the deck be-
cause of the maximum principal stress. specimens had more oblique cracks, which might have been caused
Fig. 16 shows the origin coordinates of the root tip and the location by the geometric shape of the root gap.
of all crack tips. In Fig. 16(a), the SP0 specimens have the smallest Similar results were obtained by related experiments in past studies.
crack angles, and the average cracking angles of the SP75 and SP100 Small-scale specimens of a rib-to-deck welded joint with no restraint on
specimens are nearly the same. The crack depth of SP75 ranged from the ribs were tested. It reported that the fatigue resistance of a 100%
1.6 mm to 4.6 mm, while SP100 had a crack depth of around penetration weld was slightly lower than that of an 80% PJP weld,
4 ~ 6 mm in the deck vertical direction, as shown in Fig. 16(b). SP100 which is consistent with the findings that a single-sided welded speci-
clearly had the largest crack length in both y and z directions. In order men tends to have a higher fatigue resistance as the weld penetration
to eliminate the interference effect of small cracks, the three longest rate is increased from 0% to 75%. An analytical study also suggested
cracks of every specimen were compared, which facilitated a compari- that an increase in the penetration rate (up to 75%) would produce a
son of the maximum crack depth within a segment of 50 mm in the lon- lower stress at the weld root, which results in a higher fatigue resistance
gitudinal direction. As shown in Fig. 17, the SP75 specimens had the [24,25]. We considered that a penetration rate in the range of 0% to 75%
smallest crack depths and largest propagation angles. The SP0 is desirable for fatigue durability of this structural detail. However,

Fig. 10. Residual stress distribution comparison. Fig. 12. Measured reference stress range of D12U6SP75.
S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122 121

Fig. 15. Depth distribution of deck/root cracks of D12U6SP0-4.

Fig. 13. Crack depths comparison of D12U6SP0-1 cross sections.

increasing the penetration rates above 75% might have an opposite


effect.

6. Conclusion

This study experimentally investigated the fatigue behavior of the


weld root in an orthotropic steel deck stiffened with U-ribs in relation
to the loading conditions and welding details. Field measurements, re-
sidual stress tests, and fatigue tests were carried out on a total of 12
full-scale specimens, which had 0%, 75%, and 100% penetration rates.
The fatigue behaviors were compared with those of other recent

Fig. 14. Comparison of crack propagation angles of D12U6SP0-1 cross sections. Fig. 16. Average values of all crack tips.
122 S. Kainuma et al. / Journal of Constructional Steel Research 119 (2016) 113–122

References
[1] Z.H. Qian, D. Abruzzese, Fatigue failure of welded connections at orthotropic bridges,
Giornata IGF Forni di Sopra (UD), 9 2009, p. 105.
[2] S. Sakai, Innovative Technologies for the Steel Bridges of Expressway in Japan, 2010.
[3] S. Abdou, W. Zhang, J.W. Fisher, Orthotropic deck fatigue investigation at Triborough
bridge, New York. Transportation research board 2003 meeting.
[4] Japan Society of Civil Engineers, Committee on steel structures. Fatigue of
orthotropic steel bridge deck, Steel Struct. (19) (2010) (In Japanese).
[5] F.B.P. DeJong, Renovation Techniques for Fatigue Cracked Orthotropic Steel Bridge
Decks, Delft University of Technology, 2007 (Doctoral dissertation).
[6] X. Ju, S.M. Choi, K. Tateishi, Analytical study on crack propagation at rib-to-deck
welded joints in orthotropic steel decks, Steel Constr. Eng. Jpn. Soc. Steel Constr.
19 (73) (2012) 85–94.
[7] C.V. Dung, E. Sasaki, K. Tajima, T. Suzuki, Investigations on the effect of weld pene-
tration on fatigue strength of rib-to-deck welded joints in orthotropic steel decks,
Int. J. Steel Struct. 12 (2014) 1–12.
[8] H.B. Sim, C.M. Uang, C. Sikorsky, Effects of fabrication procedures on fatigue resis-
tance of welded joints in steel orthotropic decks, J. Bridg. Eng. 14 (5) (2009)
366–373.
[9] S. Inokuchi, Wheel trucking test for weld of U-shaped rib and deck plate in the
orthotropic deck, Proceedings of the 23th U.S.-Japan Bridge Engineering Workshop,
11 2007, pp. 325–336.
[10] S. Ya, K. Yamada, T. Ishikawa, Fatigue durability of trough rib to deck plate welded
detail of some orthotropic steel decks, J. Struct. Eng. Jpn. Soc. Civ. Eng. 56 (A)
(2010) 77–90.
[11] S. Ya, K. Yamada, Fatigue durability evaluation of trough to deck plate welded joint
of orthotropic steel deck, J. Jpn .Soc. Civ. Eng A 64 (3) (2008) 603–616.
[12] H.B. Sim, C.M. Uang, Stress analyses and parametric study on full-scale fatigue tests
of rib-to-deck welded joints in steel orthotropic decks, J. Bridg. Eng. 17 (2012) 765-
Fig. 17. Comparison of angles and depths of crack tips. 733.
[13] S. Inokuchi, D. Uchida, A. Kawabata, T. Tamakoshi, Influence of asphalt pavement
failure on local stress of orthotropic steel decks, Steel Constr. Eng. Jpn. Soc. Steel
Constr. 15 (59) (2008) 75–86 (In Japanese).
[14] S. Kainuma, S. Onoue, K. Miura, S. Inokuchi, A. Kawabata, D. Uchida, Development of
studies. Based on the experimental data and parametric analysis, the an experimental system for fatigue-cracking from weld roots between deck plate
and u-rib in orthotropic steel decks, J. Jpn. Soc. Civ. Eng. A 64 (2) (2008) 297–302
main conclusions of the study are as follows:
(In Japanese).
[15] S. Inokuchi, S. Kainuma, A. Kawabata, D. Uchida, Field measurement and develop-
- Root cracks tend to initiate easily in the basic stress range − 160 to
ment of an experimental system for fatigue-cracking from weld roots between
20 MPa under double tire loading. The tested specimens could be deck plate and U-rib in orthotropic steel decks, Proceeding of 2nd International
divided into three types: 1) Specimens having only root crack; Orthotropic Bridge Conference, Sacramento Section, American Society of Civil
2) Specimens having both toe crack and root crack and usually Engineers, Sacramento, California, USA 2008, pp. 345–357.
[16] Japanese Standards Association, Japanese Industrial Standard G3106: Rolled steels
at N ar under the condition of high stress ratio; 3) Specimens having for welded structure, 2015.
no cracks. [17] C. Miki, Fatigue damage in orthotropic steel bridge decks and retrofit works, Int. J.
- The tensile residual stress existed around the weld root tip. The basic Steel Struct. 6 (4) (2006) 255–267.
[18] S.J. Maddox, Fatigue Strength of Welded Structures, Woodhead publishing, 1991.
trends of residual stress of specimens with different penetration [19] C. Acevedo, A. Nussbaumer, J.M. Drezet, Evaluation of residual welding stresses and
rates were similar in the vertical direction of the deck, and the 0% fatigue crack behavior in tubular K-joints in compression, Stahlbau 80 (7) (2011)
penetration specimens had the smallest residual stress according 483–491.
[20] S. Inokuchi, S. Kainuma, D. Uchida, D. Shiro, Influence of press reforming in fabrica-
to a theoretical analysis of cracking initiation point. tion process on stress properties of welded joint between deck plate and U-shaped
- Based on the fatigue test results, both the tensile stress and stress rib in orthotropic steel decks, Steel Constr. Eng. Jpn. Soc. Steel Constr. 19 (73) (2013)
range could affect the root crack initiation. Based on the residual 1–8 (In Japanese).
[21] N. Tebedge, G. Alpsten, L. Tall, Residual-stress measurement by the sectioning meth-
stress test results, the tensile stress would be the effective stress od, Exp. Mech 13 (2) (1973) 88–96.
for root crack propagation because the residual stress of the crack [22] S. Kainuma, D. Takamatsu, Fatigue behavior of fillet welded cruciform joint under
tip would be released after crack initiation. The root crack appeared compressive cyclic stress, J. Constr. Steel Jpn. Soc. Steel Constr. 8 (2000) 723–730
(In Japanese).
to close after the residual stress was released, which finally stopped
[23] S. Kainuma, T. Mori, A study on fatigue crack initiation point of load-carrying fillet
the crack propagation. welded cruciform joints, Int. J. Fatigue 30 (9) (2008) 1669–1677.
- Specimens with a 0% penetration rate had more oblique root cracks [24] S. Ya, K. Yamada, T. Ishikawa, Fatigue evaluation of rib-to-deck welded joints of
based on a crack morphology comparison. In general, a penetration orthotropic steel bridge deck, J. Bridg. Eng. 16 (4) (2011) 492–499.
[25] T. Mori, Influence of weld penetration of fatigue strength of single-sided fillet
rate in the range of 0% to 75% was beneficial for fatigue durability welded joints, Steel Constr. Eng. Jpn. Soc. Steel Constr. 10 (40) (2003) 9–15
of this structural detail. (In Japanese).

You might also like