Multi Scale Reliability and Serviceability Assessment of in Service PDF
Multi Scale Reliability and Serviceability Assessment of in Service PDF
LU • NOORI
FOR THE SUSTAINABLE STRUCTURAL
ENGINEERING
Serviceability Assessment of SYSTEMS COLLECTION
LIBRARY In-service Long-span Bridges Mohammad Noori, Editor
Create your own
Naiwei Lu • Mohammad Noori
Customized Content
Bundle—the more With the development in global economic and transportation
books you buy, engineering, the traffic loads on brides have been g
rowing
Multi-scale
heavy traffic volumes and simultaneous truck loads on the
THE CONTENT bridge deck, and thus the safety and serviceability of the bridge
• Manufacturing deserves investigation. In this book, a multiscale reliability
Engineering
• Mechanical
method is presented for the safety assessment of long-span
bridges. The multiscale failure condition of stiffness girders Reliability and
Serviceability
& Chemical is the first-passage criteria for the large-scale model and the
Engineering fatigue damage criteria for the small-scale model.
• Materials Science
& Engineering
• Civil &
Environmental
It is the objective of this book to provide a more in-depth
understanding of the vehicle-bridge interaction from the
random vibration perspective. This book is suitable for
Assessment
Engineering
• Advanced Energy
adoption as a text book or a reference book in an advanced
structural reliability analysis course. Furthermore, this book also
of In-service
Long-span
Technologies provides a theoretical foundation for better understanding of
the safety assessment, operation management, maintenance
THE TERMS and reinforcement for long-span bridges and motivates further
• Perpetual access for
a one time fee
• No subscriptions or
research and development for more advanced reliability and
serviceability assessment techniques for long-span bridges. Bridges
access fees Naiwei Lu is a lecturer of civil engineering at Changsha
• Unlimited University of Science and Technology.
concurrent usage
• Downloadable PDFs Mohammad Noori is professor of mechanical engineering and
• Free MARC records ASME fellow in the department of mechanical engineering,
ISBN: 978-1-94708-338-7
MULTI-SCALE
RELIABILITY AND
SERVICEABILITY
ASSESSMENT OF
IN-SERVICE LONG-
SPAN BRIDGES
MULTI-SCALE
RELIABILITY AND
SERVICEABILITY
ASSESSMENT OF
IN-SERVICE LONG-
SPAN BRIDGES
10 9 8 7 6 5 4 3 2 1
KEYWORDS
List of Figures xi
Introduction
1.1 RESEARCH SIGNIFICANCE
1.2 STATE-OF-THE-ART REVIEW
To the best of the authors’ knowledge, very limited and scarce research
work has been reported in the literature on the probabilistic dynamic anal-
ysis of long-span bridges subjected to combined stochastic traffic and
wind excitations, which is extremely important for the safety of long-span
bridges. Furthermore, most dynamic analyses reported in this area have
focused on time domain analysis, because of the nature of time-varying
differential equations in the interaction system, while very limited devel-
opments have been done in the frequency domain. However, incorporat-
ing the random vibration in the aforementioned coupled system, which
requires and necessitates the use of spectral analysis, is more important
and results in more valuable information in the frequency domain.
First-passage principle and fatigue damage principle are two main assump-
tions for structural dynamic reliability evaluation. The first-passage reli-
ability can be described as estimating the probability that a random process
exceeds a prescribed threshold during an interval of time. Knowledge of
this probability is essential for estimating the reliability of a structural
dynamic system whose response is a stochastic process. The fatigue dam-
age criterion should be adopted for the accumulated fatigue damage at
the critical regions of a structure such as connections and joints. This
topic will be discussed in more details in the next section. The classical
first-passage criterion was originally proposed by Rice [24] based on the
random vibration and extreme value distribution theory. Mathematical
formulations for the number of times that the structural responses cross
the limits were also established by Rice. A well-known crossing process
is the Rayleigh distribution as the extreme value distribution of a nar-
row-banded Gaussian stochastic process. The Poisson’s assumption and
the Vanmarcke’s assumption were widely used for general stochastic pro-
cesses in engineering structure [25]. However, these assumptions come
more from intuition or empirical approaches, rather than from theoreti-
cal basis. For this reason, several researchers developed approaches for
the improvement of the assumptions that were based on empirical work.
A joint first-passage probability method was proposed to evaluate the
reliability of linear engineering systems composed of several interdepen-
dent components by Song and Der [26]. For nonlinear dynamic systems,
Cai and Lin [27], investigated the first passage problem using stochastic
averaging. Noori, et al. [28] introduced the first-passage study of a highly
Introduction • 7
REFERENCES
[1] Xie, H., Y. Wang, H. Wu, and Z. Li. 2014. “Condition Assessment of
Existing RC Highway Bridges in China Based on SIE2011.” Journal
of Bridge Engineering 19, no. 12, 04014053. doi:10.1061/(ASCE)
BE.1943-5592.0000633
[2] Frangopol, D.M., and M. Soliman. 2015. “Life-cycle of Structural
Systems: Recent Achievements and Future Directions[J].” Structure and
Infrastructure Engineering 12, no. 1, 1–20. https://doi.org/10.1080/15732
479.2014.999794
Introduction • 9
[3] Han, W., J. Wu, C.S. Cai, and S. Chen. 2014. “Characteristics and Dynamic
Impact of Overloaded Extra Heavy Trucks on Typical Highway Bridges[J].”
Journal of Bridge Engineering 20, no. 2, p. 05014011.
[4] Wardhana, K., and F.C. Hadipriono. 2003. “Analysis of Recent Bridge
Failures in the United States[J].” Journal of Performance of Constructed
Facilities 17, no. 3, pp. 144–50.
[5] Deng, L., W. Wang, and Y. Yu. 2015. “State-of-the-Art Review on the
Causes and Mechanisms of Bridge Collapse[J].” Journal of Performance of
Constructed Facilities 30, no. 2, p. 04015005.
[6] Xing, C., H. Wang, A. Li, and Y. Xu. 2013. “Study on Wind-induced
Vibration Control of a Long-span Cable-stayed Bridge Using TMD-type
Counterweight [J].” Journal of Bridge Engineering 19, no. 1, pp. 141–48.
[7] Brownjohn, J.M.W. 1997. “Vibration Characteristics of a Suspension
Footbridge[J].” Journal of Sound and Vibration 202, no. 1, pp. 29–46.
[8] Cai, C.S., J. Hu, S. Chen, Y. Han, W. Zhang, and X. Kong. 2015. “A Coupled
Wind-vehicle-bridge System and its Applications: A Review[J].” Wind and
Structures 20, no. 2, pp. 117–42.
[9] Li, Y., P. Hu, C.S. Cai, M. Zhang, and S. Qiang. 2012. “Wind Tunnel Study
of a Sudden Change of Train Wind Loads Due to the Wind Shielding Effects
of Bridge Towers and Passing Trains[J].” Journal of Engineering Mechanics
139, no. 9, pp. 1249–59.
[10] Aktan, A.E., D.N. Farhey, D.L. Brown, V. Dalal, A.J. Helmicki, V.J. Hunt,
and S.J. Shelley. 1996. “Condition Assessment for Bridge Management[J].”
Journal of Infrastructure Systems 2, no. 3, pp. 108–17.
[11] Deng, L., Y. Yu, Q. Zou, and C. Cai. 2015. “State-of-the-Art Review of
Dynamic Impact Factors of Highway Bridges.” Journal of Bridge Engineering
20, no. 5, 04014080. doi:10.1061/(ASCE)BE.1943-5592.0000672
[12] Chen, S.R., and J. Wu. 2009. “Dynamic Performance Simulation of Long-
span Bridge Under Combined Loads of Stochastic Traffic and Wind[J].”
Journal of Bridge Engineering 15, no. 3, pp. 219–30.
[13] Hao, S. 2009. “I-35W Bridge Collapse[J].” Journal of Bridge Engineering
15, no. 5, pp. 608–14.
[14] Morales-Nápoles, O., and R.D.J.M. Steenbergen. 2015. “Large-Scale Hybrid
Bayesian Network for Traffic Load Modeling from Weigh-in-Motion System
Data[J].” Journal of Bridge Engineering 20, no. 1, pp. 591–99.
[15] Schadschneider, A. 2002. “Traffic Flow: A Statistical Physics Point of
View[J].” Physica A: Statistical Mechanics and its Applications 313, no. 1,
pp. 153–87.
[16] Enright, B., and E.J. O’Brien. 2013. “Monte Carlo Simulation of Extreme
Traffic Loading on Short and Medium Span Bridges[J].” Structure and
Infrastructure Engineering 9, no. 12, pp. 1267–82.
[17] Chen, S.R., and J. Wu. 2011. “Modeling Stochastic Live Load for Long-
span Bridge Based on Microscopic Traffic Flow Simulation[J].” Computers
& Structures 89, no. 9, pp. 813–24.
10 • BRIDGE RELIABILITY AND SERVICEABILITY
[18] Guo, W.H., and Y.L. Xu. 2001. “Fully Computerized Approach to Study
Cable-stayed Bridge–Vehicle Interaction[J].” Journal of Sound and Vibration
248, no. 4, pp. 745–61.
[19] Ahmari, S., M. Yang, and H. Zhong. 2015. “Dynamic Interaction Between
Vehicle and Bridge Deck Subjected to Support Settlement[J].” Engineering
Structures 84, pp. 172–83.
[20] Chen, S.R., and J. Wu. 2010. “Dynamic Performance Simulation of Long-
span Bridge Under Combined Loads of Stochastic Traffic and Wind[J].”
Journal of Bridge Engineering 15, no. 3, pp. 219–30.
[21] Cai, C.S., and S.R. Chen. 2004. “Framework of Vehicle–bridge–wind
Dynamic Analysis[J].” Journal of Wind Engineering and Industrial
Aerodynamics 92, no. 7, pp. 579–607.
[22] Chen, S.R., and C.S. Cai. 2007. “Equivalent Wheel Load Approach for
Slender Cable-stayed Bridge Fatigue Assessment Under Traffic and Wind:
Feasibility Study[J].” Journal of Bridge Engineering 12, no. 6, pp. 755–64.
[23] Wu, J., and S.R. Chen. 2011. “Probabilistic Dynamic Behavior of a Long-
span Bridge Under Extreme Events[J].” Engineering Structures 33, no. 5,
pp. 1657–65.
[24] Rice, S.O. 1945. “Mathematical Analysis of Random Noise[J].” Bell System
Technical Journal 24, no. 1, pp. 46–156.
[25] Vanmarcke, E.H. 1975. “On the Distribution of the First-Passage Time for
Normal Stationary Random Process[J].” Journal of Application Mechanics
42, pp. 215–20.
[26] Song, J., and A. Der Kiureghian. 2006. “Joint First-Passage Probability and
Reliability of Systems Under Stochastic Excitation.” Journal of Engineering
Mechanics 132, no. 1, pp. 65–77.
[27] Cai, G.Q., and Y.K. Lin. 1994. “On Statistics of First-passage Failure[J].”
Journal of Applied Mechanics 61, no. 1, pp. 93–99.
[28] Noori, M., M. Dimentberg, Z. Hou, R. Christodoulidou, and A. Alexandrou.
1995. “First-passage Study and Stationary Response Analysis of a
BWB Hysteresis Model Using Quasi-conservative Stochastic Averaging
Method[J].” Probabilistic Engineering Mechanics 10, no. 3, pp. 161–70.
[29] Macke, M., and C. Bucher. 2003. “Importance Sampling for Randomly
Excited Dynamical Systems[J].” Journal of Sound and Vibration 268, no.
2, pp. 269–90.
[30] Chen, J.B., and J. Li. 2007. “The Extreme Value Distribution and Dynamic
Reliability Analysis of Nonlinear Structures with Uncertain Parameters[J].”
Structural Safety 29, no. 2, pp. 77–93.
[31] Park, Y.J., and A.H.S. Ang. 1985. “Mechanistic Seismic Damage Model
for Reinforced Concrete[J].” Journal of Structural Engineering 111, no. 4,
pp. 722–39.
[32] Zhang, Z.C., J.H. Lin, Y.H. Zhang, Y. Zhao, W.P. Howson, and F.W. Williams.
2010. “Non-stationary Random Vibration Analysis for Train–bridge Systems
Subjected to Horizontal Earthquakes[J].” Engineering Structures 32, no. 11,
pp. 3571–82.
Introduction • 11
[33] Balafas, K., and A.S. Kiremidjian. 2015. “Reliability Assessment of the
Rotation Algorithm for Earthquake Damage Estimation[J].” Structure and
Infrastructure Engineering 11, no. 1, pp. 51–62.
[34] Xiang, T., R. Zhao, and T. Xu. 2007. “Reliability Evaluation of Vehicle–
bridge Dynamic Interaction[J].” Journal of structural Engineering 133,
no. 8, pp. 1092–99.
[35] Guo, T., and Y.W. Chen. 2013. “Fatigue Reliability Analysis of Steel
Bridge Details Based on Field-monitored Data and Linear Elastic Fracture
Mechanics[J].” Structure and Infrastructure Engineering 9, no. 5,
pp. 496–505.
[36] Chen, N.Z., G. Wang, and C. Guedes Soares. 2011. “Palmgren–Miner’s
Rule and Fracture Mechanics-Based Inspection Planning[J].” Engineering
Fracture Mechanics 78, no. 18, pp. 3166–82.
[37] Sim, H.B., and C.M. Uang. 2012. “Stress Analyses and Parametric Study on
Full-scale Fatigue Tests of Rib-to-deck Welded Joints in Steel Orthotropic
Decks[J].” Journal of Bridge Engineering 17, no. 5, pp. 765–73.
[38] Chen, Z.W., Y.L. Xu, and X.M. Wang. 2011. “SHMS-based Fatigue
Reliability Analysis of Multiloading Suspension Bridges[J].” Journal of
Structural Engineering 138, no. 3, pp. 299–307.
[39] Zhang, W., C.S. Cai, and F. Pan. 2012. “Fatigue Reliability Assessment
for Long-span Bridges Under Combined Dynamic Loads from Winds and
Vehicles[J].” Journal of Bridge Engineering 18, no. 8, pp. 735–47.
[40] Zhang, W., C.S. Cai, F. Pan, and Y. Zhang. 2014. “Fatigue Life Estimation
of Existing Bridges Under Vehicle and Non-stationary Hurricane Wind[J].”
Journal of Wind Engineering and Industrial Aerodynamics 133, pp. 135–45.
[41] Zhang, W., C.S. Cai, and F. Pan. 2013. “Nonlinear Fatigue Damage
Assessment of Existing Bridges Considering Progressively Deteriorated
Road Conditions[J].” Engineering Structures 56, pp. 1922–32.
[42] Guo, T., D.M. Frangopol, and Y. Chen. 2012. “Fatigue Reliability Assessment
of Steel Bridge Details Integrating Weigh-in-motion Data and Probabilistic
Finite Element Analysis[J].” Computers & Structures 112, pp. 245–57.
[43] Wang, Y., Z.X. Li, and A.Q. Li. 2010. “Combined Use of SHMS and Finite
Element Strain Data for Assessing the Fatigue Reliability Index of Girder
Components in Long-span Cable-stayed Bridge[J].” Theoretical and Applied
Fracture Mechanics 54, no. 2, pp. 127–36.
[44] Deng, Y., Y.L. Ding, A.Q. Li, and G. Zhou. 2011. “Fatigue Reliability
Assessment for Bridge Welded Details Using Long-term Monitoring
Data[J].” Science China Technological Sciences 54, no. 12, pp. 3371–81.
[45] Kwon, K., and D.M. Frangopol. 2011. “Bridge Fatigue Assessment and
Management Using Reliability-based Crack Growth and Probability of
Detection Models[J].” Probabilistic Engineering Mechanics 26, no. 3,
pp. 471–80.
CHAPTER 2
Fatigue Reliability
Assessment of Welded
Steel Bridge Decks Under
Stochastic Truck Loads
2.1 INTRODUCTION
6.0 m
4.27m 9.14m
(a)
1.2m 1.2m
1.8m
0.4m
0.4m
(b)
0.02
Measured data
GMM
0.015
Probability density
0.01
0.005
0
0 500 1000 1500
GVW(kN)
(a)
0.12
Measured data
GMM
Probability density
0.08
0.04
0
100 200 300
AW64(kN)
(b)
slow lane are shown in Figure 2.3. The GVW, instead of the axle weight,
was used in Figure 2.3 for more clear and thorough presentation of the
results.
As shown in Figure 2.3, each dot refers to a specific truck with dif-
ferent labeling marks, x-axis shows the arrival time, and y-axis shows the
individual GVW. It is observed that each truck is different due to its spe-
cific characteristics. However, all trucks follow a relatively similar corre-
sponding probability distribution. In addition, the distributions of vehicle
configuration in both fast lane and slow lane are obviously different,
20 • BRIDGE RELIABILITY AND SERVICEABILITY
1200
V1
V2
900 V3
V4
V5
GVW (kN)
600 V6
300
0
0 0.5 1 1.5 2 2.5
Arrival time (h)
(a)
1200
V1
V2
V3
900
V4
V5
GVW (kN)
V6
600
300
0
0 0.5 1 1.5 2 2.5
Arrival time (h)
(b)
Figure 2.3. Simulated stochastic fatigue truck loads in: (a) slow lane;
(b) fast lane.
2.3 COMPUTATIONAL FRAMEWORK OF
PROBABILISTIC MODELING
For the purpose of applying the stochastic traffic truck load model to prob-
abilistic modeling of fatigue stress in steel bridge decks, an efficient com-
putational framework should be developed as a connection between the
stochastic load model and the structural effect model. The innovation of
the work presented in this book is the application of stochastic truck model,
presented earlier, instead of the commonly used and the typical truck
model. However, this will be a computationally time-consuming problem
for time history analysis taking into account each truck passage, because
a large number of finite element analysis runs are required to simulate the
fatigue behavior of welded details accurately. Therefore, improving the
computational efficiency, by not ignoring the importance of the require-
ments for the computational accuracy, is a critical step for the application
of the stochastic fatigue truck load in the study presented in this book.
In order to address the exhaustive computation time problem, a
machine learning algorithm was employed by integrating a uniform
design (UD) and a support vector regression (SVR) approach. This pro-
posed learning machine is utilized to approximate the response surface
between the vehicle axles and the equivalent stress ranges for a certain
vehicle configuration. The flowchart summarizing the entire procedure
is depicted in Figure 2.4. There are two main procedures, as illustrated
Hot spot stress range simulation: Estimating Δσre and Ned in the large sample:
compute the stress time history based on FEA estimate the Δσre and corresponding number of
and extract the equivalent stress ranges, daily stress cycles, Ned, for the individual truck
Δσre, of the welded connections under in the large sample utilizing the approximated
individual vehicle load in the small samples. learning machine.
and the corresponding number of cycles can be calculated by the rain flow
counting method. The multi-amplitude stress cycles can be equivalent to a
constant-amplitude stress cycle according to the fatigue damage accumu-
lation rule that is illustrated in the next section.
The most essential aspect for the development of the deterministic
analysis in the research presented in this book is the approximating func-
tion linking the vehicle weight and the equivalent fatigue stress ranges.
A simple illustration of the SVR formulations is presented in the follow-
ing. In general, the purpose of SVR is to find a relationship between the
input space and the output space with the given training dataset, {(xi, yi), i
= 1,…,N}, where the xi and yi are the ith input data and output data, respec-
tively and N is the number of total data. The general form of the SVR is
written as [28]
l
f ( x) = ∑ ai È ( x, xi ) + b (2.1)
i =1
2.3.3 PROBABILISTIC MODELING
As the fatigue stress ranges are calculated in the deterministic finite ele-
ment analysis, the next step is the probabilistic modeling of these fatigue
24 • BRIDGE RELIABILITY AND SERVICEABILITY
stress ranges. The probabilistic modeling results can provide the neces-
sary statistics for the reliability analysis. It can be observed from the sta-
tistics of traffic data that the parameters in the truck load model follow
various types of probabilistic distributions. As a result, the probabilistic
density of the fatigue stress range may not follow a certain single distribu-
tion. Regarding the random variables in the stochastic fatigue truck load,
the probabilistic distribution of the axle weight is the critical factor that
directly influences the probabilistic distribution of fatigue stress ranges.
Since the axle weight shown in Figure 2.2 follows a multi-peak distri-
bution, an appropriate probabilistic density function should be considered.
The GMM is also used herein to approximate the probability density of
fatigue stress. The GMM is part of the finite mixture distributions that
are commonly employed for modeling complex probability distributions.
They enable the statistical modeling of random variables with multimodal
behaviors. The basic structure of finite mixture distributions for indepen-
dent scalar y can be expressed as [34]
c
f ( y | c, w, q ) = ∑ wi f i ( y | qi ) (2.2)
i =1
c
∑w
i =1
i = 1 (2.3)
c
1 1 ( y − mi )
f ( y | c, w, q ) = ∑ wi exp − (2.4)
2 si
2
i =1 2À
where μi and σi are the mean value and the standard deviation of the ith
normal mixture parameter. It is observed that the GMM was primly used
for the probabilistic modeling of vehicle weight, and then was used for
the probabilistic modeling of structural fatigue stress range. In addition,
there is a relationship between the vehicle weight and the structural fatigue
stress range. Therefore, the GMM can provide a reliable connection for
the monitored traffic data and the probabilistic modeling of structural
fatigue stress range.
FATIGUE RELIABILITY OF STEEL BRIDGES • 25
ni ∆si3 n j ∆s 5j
D=∑ +∑ (2.7)
i KC j KD
where D is the fatigue damage accumulation, Δσi is the ith fatigue stress
range that Δσi≥ΔσD, Δσj is the jth fatigue stress range that Δσj<ΔσD, ni and
nj are the corresponding number of cycles for Δσi and Δσj, respectively.
With the equivalent fatigue damage accumulation rule, the equivalent
fatigue stress range and corresponding number of cycles are expressed as
ni ∆si3 n j ∆s 5j
∑i K + ∑j K
C D
∆sre5 = (2.8)
N axle / K D
where Δσre is the equivalent fatigue stress range, and Naxle is the corre-
sponding number of stress cycles that equal to the axle number of a truck.
Herein, the variable amplitude stresses can be transformed to constraint
amplitude stresses. Subsequently, the S-N curves in the Eurocode 3 spec-
ification can be used to calculate the fatigue damage.
The previous illustration focuses on fatigue damage expressions
caused by fatigue stress cycles due to an individual truck passage. In prac-
tice, Δσre mainly depends on the axle weight that is random in nature and
should be treated as a random variable. In addition, however, the rela-
tionship between fatigue stress cycles and passing trucks can also affect
the fatigue damage. Thus, additional parameters should be considered in
the limit state function of the fatigue damage. These include the transverse
distribution factor of truck axles, the average daily truck traffic (ADTT),
and the growth factor of traffic volume and the axle weight. Subsequently,
the limit state function can be expressed as
FATIGUE RELIABILITY OF STEEL BRIDGES • 27
n
g n (X ) = D∆ − ∑ Di (X )
i =1
n
= D∆ − 365 N ADTT ∆sre5 N ed w / K D ⋅ ∑ (1 + ia )(1 + ib)5 (2.9)
i =1
2.5 CASE STUDY
Dimensions of the cross section and the U-rib are shown in Figure 2.5.
The half-segment finite element model established by shell63 elements
in ANSYS is shown in Figure 2.6. The deck and the ribs were meshed
with the quadrilateral elements, while the longitudinal stiffening plates,
the diaphragm plates, and the web plates were meshed by triangular ele-
ments. Since the pavement elements have not been considered in the
finite element model, a spreading angle of 45 [27] for a vertical uniformly
distributed wheel load was applied on the bridge deck. For instance, the
thickness of pavement in this case study is 6.7cm, and the load areas are
30cm × 20m and 60cm × 20cm for the front and back wheels, respec-
tively. Therefore, with consideration of the pavement, the corresponding
28 • BRIDGE RELIABILITY AND SERVICEABILITY
13.8m 13.8m
Sidewalk 1st lane 2nd lane 3rd lane 4th lane
2% 2%
3m
Cracks
30cm Bridge
deck
U-rib
28c
Joint I:
m
Rib-to-deck joint
17cm
Joint II:
Butt joint of U-rib
(b)
Slow lane
Fast lane
4
Joint I in model I
Joint I in model II
2 Joint II in model I
Joint II in model II
0
Stress (MPa)
-2
-4
Diaphragm plate
-6
-8
0 2 4 6 8 10 12
Longitudinal location in a segment (m)
the three-segment model. Even though the segment model has a simi-
lar fatigue behavior with the global model [21], a global model is still
important for a more detail comparison. Nevertheless, this verifies that
the vehicle configuration is sensitive to the fatigue stress, while the vehi-
cle spacing is insensitive.
In order to observe stress–time histories of the welded joints under
truck loads, two V6 trucks with a maximum and a minimum gross weight,
respectively, were added to the finite element model with a constant speed
of 20 m/s. With the transient analysis of the finite element model, the
results are shown in Figure 2.8, where the load case is critical for the
corresponding fatigue stress of the welded joints, and the σmin and σmax
denote the fatigue stress under V6 truck loads with minimum and maxi-
mum GVW, respectively. As observed from the stress–time histories, the
stress amplitudes are time varying. This clearly indicates the necessity of
the equivalent stress range simplification in Equation (2.8). In addition,
the shape of the stress history is associated with the vehicle configuration,
which also demonstrates the importance of the vehicle configuration in
stochastic fatigue truck model. Based on these time–stress histories, the
fatigue stress ranges and the corresponding number of cycles can be cal-
culated by utilizing the rain flow counting method.
30 • BRIDGE RELIABILITY AND SERVICEABILITY
20
σmin
10
σmax
0
Stress (MPa)
–10
–20
–40
–50
0 0.2 0.4 0.6 0.8 1 1.2 1.4
(a)
Time (s)
50
40 σmin
30 σmax
20
Stress (MPa)
10
0
Load case
–10
–20
0 0.2 0.4 0.6 0.8 1 1.2 1.4
(b)
Time (s)
Figure 2.8. Stress–time histories under the V6 truck loads for the welded
joints of: (a) rib-to-deck joint; (b) butt joint of U-rib.
2.5.2 PROBABILISTIC ANALYSIS
U10 and U20, standing for the total number of 10 and 20 uniform distrib-
uted samples, respectively, were selected for approximating the response
surface between the structural equivalent fatigue stress range and the vehi-
cle axle weight. After performing the necessary finite element analyses
for 10 and 20 times, the response surfaces were obtained as shown in
Figure 2.9(a) and Figure 2.9(b), respectively. As shown in Figure 2.9, both
training samples are uniformly distributed in the coordinate space, and the
approximated response surfaces are close to the training data. In addition,
the approximated response surfaces are nonlinear, and the SVR approach
is appropriate for the approximation, since its kennel function is nonlinear.
80
60
40
20
–20
500
400 300
300
200 200
AW22 (kN) 100 100
0 0 Negative AW21 (kN)
(a)
80
60
40
20
0
500
400 300
300
200 200
AW22 (kN) 100 100
0 0 AW21 (kN)
(b)
Figure 2.9. Response surfaces for the 2rd type of vehicle with the training
sample of: (a) U10; (b) U20.
32 • BRIDGE RELIABILITY AND SERVICEABILITY
However, when compared with Figure 2.9(a) and Figure 2.9(b), it is easy
to find out that the regions in the corner for both response surfaces are
different, where Δσre with U10 data are negative, while Δσre with U20 data
are positive. The reason for the invalidity of the response surface with
U10 training samples is the lack of training data in the margins. Figure
2.10 shows the comparison between the SVR predictions and the FEA
results, where AW21 was supposed to be 150 kN, and AW22 was supposed
to change from 50 to 500 kN. It can be observed that the SVR predictions
with U20 is in agreement with the FEA results, while two terminals of the
SVR predictions with U10 have a certain difference with the FEA results.
The analysis results indicate that a sufficient number of training samples is
a guarantee for the accuracy of SVR prediction.
With the substitution SVR model, the stress ranges under the
large-sample in the stochastic fatigue truck load model will be easy to cal-
culate. Herein, 30-day stochastic truck data were used for the probability
density modeling for the welded joints in the slow lane and the fast lane.
The results are shown in Figure 2.11 and the corresponding parameters
of the estimated PDFs are shown in Table 2.3. It can be concluded from
Figure 2.11 that the probability density curves of the stress ranges have
multi-peaks and are well fitted with the GMMs rather than the normal
distribution function. In addition, more high-amplitude stress cycles are
included in the fatigue stress spectrum of the welded joints in the slow lane,
while more low-amplitude stress cycles are included in the welded joints
in the fast lane. Note that the shape of the PDFs in Figure 2.11 is different
70
FEA
65 SVR with U10
SVR with U20
Equivalent stress range (MPa)
60
55
50
45
40
35
30
50 100 150 200 250 300 350 400 450 500
AW22 (kN)
x 10-3
5
Simulated data
GMM
4 Normal distribution
Probability density
3
0
5 10 15 20 25
Stress range (MPa)
(a)
x 10-3
7
Simulated data
6 GMM
Normal distribution
5
Probability density
0
5 10 15 20 25
Stress range (MPa)
(b)
from that calculated by Ye et al. [37] utilizing the SHM data, but they are
similar to the PDFs given by Xia et al. [23]. This result is due to two sim-
plified preconditions in the framework. The first precondition is the vehicle
filters process that vehicle weight less than 30 kN was cut off. The second
precondition is the equivalent stress process that variable amplitude stress
cycles were equivalent to be constant amplitude stress cycles.
34 • BRIDGE RELIABILITY AND SERVICEABILITY
0.004
Simulated data
Normal distribution
0.003
Probability density
m=7927
0.002 s=984
0.001
0
7600 7800 8000 8200 8400
Number of daily cycles
(a)
0.004
Simulated data
Normal distribution
0.003
Probability density
μ=13684
0.002 σ=1130
0.001
0
1.32 1.34 1.36 1.38 1.4
Number of daily cycles 4
x 10
(b)
Figure 2.12. PDFs of the equivalent number of daily cycles for: (a) slow
lane; (b) fast lane.
Table 2.4. Statistics of the random variables in the limit state function
Mean
Variables Description Distribution value COV
DΔ Critical fatigue damage Lognormal 1.0 0.3
KD Coefficient of fatigue Lognormal 3.47×1014 0.34
strength
w Coefficient of transversal Normal 0.8 1
load distribution of tires
Δσre Equivalent stress range GMM See Table 2.3 and
Figure 2.11
Nde Number of daily cycles Normal See Figure 2.12
are shown in Figure 2.12. The coefficient for the transverse distribution
of vehicle tires on the bridge deck w was assumed to follow a normal
distribution with a mean value of 0.8 and a COV of 1. In conclusion, the
statistics of the variables in the limit state function are shown in Table 2.4.
Reliability index 4
Slow lane
Fast lane
2
0 20 40 60 80 100
Service time (year)
are shown in Figure 2.14(a). Similarly, if the linear growth factor of the
vehicle weight is assumed to be 0%, 0.2%, 0.4%, and 0.6%, the corre-
sponding reliability indices are shown in Figure 2.14(b). It is observed
from Figure 2.14 that the reliability indices decrease noticeably during the
service time, especially for the case of considering the growth of vehicle
weight. For instance, while the growth factor of the traffic volume is 3%,
the corresponding reliability index in the 100th year decreases to 1.92.
When the growth factor of the vehicle weight is 0.6%, the corresponding
reliability index decreases to 1.35.
The fatigue reliability assessment can also be used to predict the
fatigue life of the steel structures under the target reliability index βtarget.
The target reliability index determines the structural cost of mainte-
nance. Helmerich et al. [30] suggest the βtarget to be in the range between
2.0 and 3.5 according to the calibration study of Eurocode 3 [36]. In the
present book, βtarget is determined to be 2.0 corresponding to the failure
probability of 2.3%. The structural fatigue life under the βtarget is shown
in Table 2.5. It is observed from Table 2.5 that the fatigue life of the
welded joints decreases significantly with the growth of traffic volume
and vehicle weight. When the linear annual growth factor of traffic vol-
ume a is greater than 3%, the corresponding fatigue life is only 97 years,
that is less than the expected design service life. When the growth factor
of vehicle weight b is greater than 0.4%, the fatigue life of the welded
details is only 94 years, which is also less than the expected design
service life.
38 • BRIDGE RELIABILITY AND SERVICEABILITY
6
a=0
a=1%
5 a=2%
a=3%
Reliability index
4
1
0 20 40 60 80 100
Service time (year)
(a)
6
b=0
b=0.2%
5
b=0.4%
b=0.6%
Reliability index
1
0 20 40 60 80 100
Service time (year)
(b)
Table 2.5. Fatigue life predication for the rib-to-deck joint taking into
account traffic growth factors
Fatigue life predication (year)
a=b a= a= a= b= b= b=
Traffic lane =0 1% 2% 3% 0.2% 0.4% 0.6%
Slow lane 195 135 108 96 134 94 72
Fast lane 297 195 161 148 186 143 120
2.7 CONCLUSIONS
REFERENCES
[1] Zhang, S., X. Shao, J. Cao, J. Cui, J. Hu, and L. Deng. 2016. “Fatigue
Performance of a Lightweight Composite Bridge Deck with Open Ribs.”
Journal of Bridge Engineering 21, no. 7, 04016039. 10.1061/(ASCE)
BE.1943-5592.0000905.
[2] Liu, M., D. Frangopol, and K. Kwon. 2010. “Fatigue Reliability Assessment
of Retrofitted Steel Bridges Integrating Monitored Data.” Structural Safety
32, no. 1, pp. 77–89.
[3] Deng, L., W. Wang, and Y. Yu. 2015. “State-of-the-art Review on the Causes
and Mechanisms of Bridge Collapse.” Journal of Performance of Constructed
Facilities 30, no. 2, 04015005. 10.1061/(ASCE)CF.1943-5509.0000731.
[4] Sim, H., and C. Uang. 2012. “Stress Analyses and Parametric Study on
Full-scale Fatigue Tests of Rib-to-deck Welded Joints in Steel Orthotropic
Decks.” Journal of Bridge Engineering, 765–73. 10.1061/(ASCE)
BE.1943-5592.0000307.
[5] Ye, X., Y. Su, and J. Han. 2014. “A State-of-the-art Review on Fatigue Life
Assessment of Steel Bridges.” Mathematical Problems in Engineering.
doi:10.1155/2014/956473.
[6] Frangopol, D., and M. Soliman. 2016. “Life-cycle of Structural Systems:
Recent Achievements and Future Directions.” Structure and Infrastructure
Engineering 12, no. 1, pp. 1–20.
[7] Chen, S., and J. Wu. 2011. “Modeling Stochastic Live Load for Long-span
Bridge Based on Microscopic Traffic Flow Simulation.” Computers &
Structures 89, no. 9, pp. 813–24.
FATIGUE RELIABILITY OF STEEL BRIDGES • 41
[8] Xu, Y., Z. Chen, and Y. Xia. 2012. “Fatigue Assessment of Multi-loading
Suspension Bridges Using Continuum Damage Model.” International
Journal of Fatigue 40, pp. 27–35.
[9] Wang, Y., Z. Li, and A. Li. 2010. “Combined Use of SHMS and Finite
Element Strain Data for Assessing the Fatigue Reliability Index of Girder
Components in Long-span Cable-stayed Bridge.” Theoretical and Applied
Fracture Mechanics 54, no. 2, pp. 127–36.
[10] Guo, T., and Y.W. Chen. 2013. “Fatigue Reliability Analysis of Steel
Bridge Details Based on Field-Monitored Data and Linear Elastic Fracture
Mechanics.” Structure and Infrastructure Engineering 9, no. 5, pp. 496–505.
[11] Ni, Y., X. Ye, and J. Ko. 2010. “Monitoring-based Fatigue Reliability
Assessment of Steel Bridges: Analytical Model and Application.” Journal
of Structural Engineering 136, no. 12, 1563–73. 10.1061/(ASCE)
ST.1943-541X.0000250.
[12] D’Angelo, L., and A. Nussbaumer. 2015. “Reliability Based Fatigue
Assessment of Existing Motorway Bridge.” Structural Safety 57, no. 7,
pp. 35–42.
[13] Laman, J., and A. Nowak. 1996. “Fatigue-load Models for Girder
Bridges.” Journal of Structural Engineering 122, no. 7, 726–733. 10.1061/
(ASCE)0733-9445.
[14] European Committee for Standardization (ECS). 2003. “Eurocode 1: Actions
on Structure—Part 2: Traffic Loads on Bridges.” EN 1991-2, Brussels,
Belgium.
[15] AASHTO. 2010. LRFD Bridge Design Specifications, 5th ed. Washington,
DC: American Association of State Highway and Transportation Officials.
[16] Zhang, W., C. Cai, and F. Pan. 2013. “Fatigue Reliability Assessment for
Long-span Bridges Under Combined Dynamic Loads from Winds and
Vehicles.” Journal of Bridge Engineering 18, no. 8, 735–47. 10.1061/
(ASCE)BE.1943-5592.0000411
[17] Zhang, W., C.S. Cai, F. Pan, and Y. Zhang. 2014. “Fatigue Life Estimation
of Existing Bridges Under Vehicle and Non-stationary Hurricane Wind.”
Journal of Wind Engineering and Industrial Aerodynamics 133, no. 10, pp.
135–45.
[18] MacDougall, C., M. Green, and S. Shillinglaw. 2006. “Fatigue Damage
of Steel Bridges Due to Dynamic Vehicle Loads.” Journal of Bridge
Engineering 11, no. 3, 320–28. 10.1061/(ASCE)1084-0702.
[19] Wang, W., L. Deng, and X. Shao. 2016. “Fatigue Design of Steel Bridges
Considering the Effect of Dynamic Vehicle Loading and Overloaded
Trucks.” Journal of Bridge Engineering, 21, no. 9, 04016048. 10.1061/
(ASCE)BE.1943-5592.0000914, 04016048.
[20] Zhang, W., and C. Cai. 2011. “Fatigue Reliability Assessment for Existing
Bridges Considering Vehicle Speed and Road Surface Conditions.” Journal
of Bridge Engineering, 443–53. 10.1061/(ASCE)BE.1943-5592.0000272.
42 • BRIDGE RELIABILITY AND SERVICEABILITY
[21] Guo, T., D. Frangopol, and Y. Chen. 2012. “Fatigue Reliability Assessment
of Steel Bridge Details Integrating Weigh-in-motion Data and Probabilistic
Finite Element Analysis.” Computers & Structures 112, pp. 245–57.
[22] Guo, T., Z. Liu, and J. Zhu. 2015. “Fatigue Reliability Assessment of
Orthotropic Steel Bridge Decks Based on Probabilistic Multi-scale Finite
Element Analysis.” Advanced Steel Construction 11, no. 3, pp. 334–46.
[23] Xia, H., Y. Ni, K. Wong, and M. Ko. 2012. “Reliability-based Condition
Assessment of in-Service Bridges Using Mixture Distribution Models.”
Computers & Structures 106–107, no. 9, pp. 204–13.
[24] Chen, W.Z., J. Xu, B.C. Yan, Z.P. Wang. 2015. “Fatigue Load Model for
Highway Bridges in Heavily Loaded Areas of China.” Adv. Steel Constr. 11,
no. 3, pp. 322–33.
[25] Chen, S., and J. Wu. 2009. “Dynamic Performance Simulation of Long-
span Bridge Under Combined Loads of Stochastic Traffic and Wind.”
Journal of Bridge Engineering 15, no. 3, 219–30. 10.1061/(ASCE)
BE.1943-5592.0000078.
[26] Han, W., J. Wu, C. Cai, and S. Chen. 2014. “Characteristics and Dynamic
Impact of Overloaded Extra Heavy Trucks on Typical Highway Bridges.”
Journal of Bridge Engineering 20, no. 2, 05014011. 10.1061/(ASCE)
BE.1943-5592.0000666.
[27] Liu, Y., N. Lu, M. Noori, and X. Yin. 2014. “System Reliability-based
Optimization for Truss Structures Using Genetic Algorithm and Neural
Network.” Int. J. of Relia. and Saf. 8, no. 1, pp. 51–69.
[28] Suykens, J., and J. Vandewalle. 1999. “Least Squares Support Vector
Machines Classifiers.” Neural Processing Letters 19, no. 3, pp. 293–300.
[29] Moura, M., E. Zio, I. Lins, and E. Droguett. 2011. “Failure and Reliability
Prediction by Support Vector Machines Regression of Time Series Data.”
Reliab. Eng. & Syst. Safe. 96, no. 11, pp. 1527–34.
[30] Helmerich, R., B. Kühn, and A. Nussbaumer. 2007. “Assessment of Existing
Steel Structures: A Guideline for Estimation of the Remaining Fatigue Life.”
Struct. and Infrastructure E. 3, no. 3, pp. 245–55.
[31] Huang, C., Y. Lee, D. Lin, and S. Huang. 2007. “Model Selection for Support
Vector Machines via Uniform Design.” Computational Statistics & Data
Analysis 52, no. 1, pp. 335–46.
[32] Deng, S., and T. Yeh. 2011. “Using Least Squares Support Vector Machines
for the Airframe Structures Manufacturing Cost Estimation.” International
Journal of Production Economics 131, no. 2, pp. 701–08.
[33] Kisi, O., and M. Cimen. 2011. “A Wavelet-Support Vector Machine
Conjunction Model for Monthly Streamflow Forecasting.” Journal of
Hydrology 399, no. 1, pp. 132–40.
[34] Deng, Y., Y. Ding, A. Li, and G. Zhou. 2011. “Fatigue Reliability Assessment
for Bridge Welded Details Using Long-term Monitoring Data.” Science
China Technological Sciences 54, no. 12, pp. 3371–81.
FATIGUE RELIABILITY OF STEEL BRIDGES • 43
[35] Jin, S., X. Qu, and D. Wang. 2011. “Assessment of Expressway Traffic Safety
Using Gaussian Mixture Model Based on Time to Collision.” International
Journal of Computational Intelligence Systems 4, no. 6, pp. 1122–30.
[36] European Committee for Standardization (ECS). 2005. “Eurocode 3: Design
of Steel Structures—Part 1–9: Fatigue.” EN1993-1-9, Brussels, Belgium.
[37] Ye, X., Y. Ni, K. Wong, and J. Ko. 2012. “Statistical Analysis of Stress
Spectra for Fatigue Life Assessment of Steel Bridges with Structural Health
Monitoring Data.” Engineering Structures 45, no. 12, pp. 166–76.
[38] Wirsching, P. 1984. “Fatigue Reliability for Offshore Structures.” Journal of
Structural Engineering 110, no. 10, pp. 2340–56.
CHAPTER 3
First-Passage Probability
of the Deflection of a
Cable-Stayed Bridge Under
Long-term Site-Specific
Traffic Loading
3.1 INTRODUCTION
The rapid growth of global economy has posed a steady traffic growth
in both the volume and the gross vehicle weight (GVW) over recent
decades. Overweight trucks have become a main cause of reducing the
service life or even collapses of existing bridges [1]. In comparison to
short- and medium-span bridges, long-span bridges suffer from higher
traffic loads and the simultaneous presence of multiple vehicles [2]. In
addition, the heavy traffic load, and the steady traffic growth may pose
an additional threat to the safety and serviceability of long-span bridges.
In general, a native design specification of traffic loading has taken
into account the probabilistic extrapolation of traffic loading as well as
the simultaneous presence of multiple vehicles. In addition, the exist-
ing bridges were designed with consideration of the uncertainties of the
traffic load and system reliability [3, 4]. However, the continuous traffic
growth and extremely overloaded trucks have changed the probability
characteristics of the traffic loading [5]. Such changes have been demon-
strated to result in a larger load effect compared to the national design
specifications. For instance, Han et al. [6] found that 4 out of 1,319 trucks
cause larger hogging moment in bridge girders, a bending moment that
produces convex bending, than, for instance, the China’s code’s [7] value.
46 • BRIDGE RELIABILITY AND SERVICEABILITY
A WIM system utilizes scales or pressure sensors embedded into the road
pavement to measure the traffic volume and the GVW of passing vehicles
[25]. In practice, WIM systems can be used for a wide range of tasks,
such as protection and management of highways and other infrastructure
investments. In the present study, WIM measurements are used for statis-
tical analysis of the traffic loading.
A bridge WIM system in Sichuan, China, was chosen herein as a pro-
totype. The WIM system has been working since the bridge was opened to
the public in May, 2012. Details of the WIM system can be found in Liu et
al. [26]. Before statistical analysis of the measurements, a filtering process
was conducted in order to identify and remove invalid records from the
database. The criteria to identify the invalid records are: (1) the GVW is
less than 30 kN; (2) the axle weight is greater than 300 kN; (3) the vehicle
length is greater than 20 m; and (4) the data are incomplete or flagged with
the system error. Overview of the filtered WIM measurements is shown in
Table 3.1. It is worth noting that the maximum GVW for a 6-axle truck is
550 kN according to the traffic laws in China [27]. However, as observed
in Table 3.1, the number of overloaded trucks is approximately 17 per day.
The overload rate of the maximum gross vehicle weight (GVW) is about
200 percent over the threshold value.
Based on these measurements, the trucks were classified into six cate-
gories. Proportion of the vehicle types are shown in Figure 3.1(a), where V1
48 • BRIDGE RELIABILITY AND SERVICEABILITY
indicates the light car, V2 to V6 indicate the 2-axle to 6-axle trucks, respec-
tively. It is observed that 39.24 percent of the entire vehicles are 3- to 6-axle
trucks, which is the main contribution to the critical traffic load scenario.
The ADTT was fitted by a normal distribution with a mean value of 2,145
and standard deviation of 424 as shown in Figure 3.1(b). The hourly traffic
density in one day is shown in Figure 3.1(c), where the traffic is divided to
free flow and busy flow. It is worth noting that the vehicle density mostly
depends on vehicle spacing, which is another factor contributed by the crit-
ical traffic load scenario. Therefore, the busy flows between 9:00 and 22:00
were used for statistical analysis of the vehicle spacing. The vehicle spac-
ing of busy flows was fitted by a lognormal distribution with a mean value
of 6.21 and standard deviation of 1.40 as shown in Figure 3.1(d).
In addition to the large-scale statistics shown in Figure 3.1, the small-
scale statistics of individual trucks were also analyzed. Taking V6 as an
example, the histogram and PDFs of the axle weight and GVW are shown
in Figure 3.2, where AW64 indicates the 4th axle of the 6-axle trucks. The
Gaussian mixture model (GMM) as shown in Figure 3.2(b) has captured
feature of the bimodal distribution. Both the large-scale and small-scale
information provide statistical parameters for the traffic load simulation.
Based on the previous demonstration, the remaining statistical anal-
yses were conducted to provide the entire probability model for the sub-
sequent stochastic traffic simulation. Most of the extreme values, which
will be used for probabilistic modeling of first-passage failure probability,
are caused by dense traffic flows. Therefore, the vehicle spacing of dense
traffic flow is effective for the traffic load simulation.
V6 15.49%
V1 35.64%
V5 4.93%
V4 10.24%
V3 8.58%
V2 26.12%
(a)
x 10–3
1.4
Monitored data
1.2 Normal distribution
1
Probability density
m=2145
0.8 s=424
0.6
0.4
0.2
0
1000 1500 2000 2500 3000 3500
(b) ADTT
0.8
Probability density
0.6
Busy flow
0.4
Free flow
0.2
0
0:00 4:00 8:00 12:00 16:00 20:00 24:00
(c) Time
1
Empirical data
Lognormal
0.8
Probability density
0.6
0.4
m=6.21
0.2 s=1.40
0
0 1000 2000 3000 4000
(d) Vehicle spacing (m)
0.12
WIM data
0.1 GMM
Probability density
a1=0.38 a2=0.62
0.08
m1=47.22 m2=132.45
0.04
0.02
0
50 100 150 200 250
(a) Axle weight AW64 (kN)
0.02
WIM data
GMM
0.015
a1=0.24 a2=0.76
PDF
0.005
0
0 500 1000 1500 2000
(b) GVW (kN)
3.3 METHODOLOGY
1000
V1
V2
800 V3
V4
GVW (kN)
600 V5
V6
400
200
0
0 10 20 30 40 50 60
(a) Arrival time (h)
1000
800
GVW (kN)
600
400
200
0
0 10 20 30 40 50 60
(b) Arrival time (min)
Load effect
x1
o t
x2
Level crossing (x>0)
Level crossing (x<0)
sure that the load effects follow a normal distribution. In this regard,
Ditlevsen [29] has demonstrated that the bridge load effect can be mod-
eled as Gaussian random process under the case of both a long influence
line and a dense traffic. Fortunately, the long-span bridge in the present
52 • BRIDGE RELIABILITY AND SERVICEABILITY
study has a long influence line. In addition, only busy traffic flows are
used for the traffic load effect analysis. Thus, the bridge load effect can
be assumed as Gaussian random process and the Rice formula is appro-
priate to be used.
Based on the preceding assumption, the mean up-crossing rate v(x)
under the condition of a threshold level is expressed as:
s′ ( x − m) 2
v( x) = exp − (3.1)
2πs 2s 2
where x is a random process, which represents the load effect in the present
study, m and σ are the mean value and the standard deviation of x, and σ¢ is
the standard deviation of the derivative of x. With consideration of the return
period, the cumulative distribution function (CDF) can be written as [30]:
1 x − m 2
F ( x) = exp −v0T exp − (3.2)
2 s
r + sr
xmax ( Rt ) = mopt
opt 2 ln (v0, optRt ) (3.3)
3.3.2 FIRST-PASSAGE PROBABILITY
t
p(a, t) ≅ 1 − A exp − ∫ v(a, t ) d t (3.5)
0
a
A = P X (0) < a = ∫ f X ( x, 0) d x (3.6)
−a
n
p(a, T ) ≅ 1 − ∑ exp −v(a, ti ) ⋅ ri ⋅ ti (3.7)
i =1
where the service period T is divided into T intervals indicated as ti, and n
is the maximum service period in year. Since the traffic density and GVW
are growing in the service period, v(a, ti) is time-variant.
No
Traffic growth model?
Yes
Life-cycle reliability evaluation
annual compounded growth rate of the traffic volume and the GVW. Since
the static influence lines are used to calculate the traffic load effect, the
computational effort can be ignored.
3.4.1 BRIDGE DETAILS
The traffic load effects were computed by the static influence lines of the
bridge. The displacement influence lines of the critical points of the bridge
FIRST-PASSAGE PROBABILITY OF THE DEFLECTION • 55
140 m
0.01
L/4
L/2
Pylon Pylon
-0.01
-0.02
-0.03
-400 -200 0 200 400
Longitudinal location (m)
0.4
L/4
Vertical displacement (m)
L/2
0.2
-0.2
-0.4
0 10 20 30 40 50 60
Time period (min)
1
L/2 data
Normalized up-crossing rate
v(x)
0.8
L/4 data
v(x)
0.6
0.4
0.2
0
0.2 0.3 0.4 0.5 0.6 0.7
Threshold displacement (m)
6
L/4 data
5 L/4 fitting
L/2 data
4 L/2 fitting
log(-log(P)
3
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7
Threshold displacement (m)
It is observed from Figures 3.10 and 3.11 that the L/4 point of the
bridge has a higher up-crossing rate and maximum load effect compared
to the L/2 point of the bridge. In addition, the tail fitting of the up-cross-
ing rate is an appropriate description of probability feature of the extreme
traffic load effect.
Based on the probability model of the traffic load effect, the extrapolation
of extreme load effects and the first-passage probability can be evaluated
with consideration of return period and threshold value, respectively. Since
the mid-span point of the bridge is the critical location for the deformation
failure, the following discussion focuses on the mid-span point. Suppose
the annual linear growth rate of both the ADTT and the GVW (RADTT and
RGVW) to be in the range between 0 and 2 percent. The maximum displace-
ment during the service period was evaluated as shown in Figure 3.12. It is
observed that the maximum displacement is 0.74 m without consideration
of the traffic growth, and the value increases with the increase of RADTT or
RGVW. However, their influences on the maximum displacement are differ-
ent. With the linear increase of RADTT the increase ratio of the maximum
displacement slows down gradually, while the increase ratio of the max-
imum displacement grows linearly with the linear increase of RGVW. For
instance, under the condition of growth rate of 2 percent, the RADTT and
RGVW will lead to the maximum of 0.87 and 0.92 m, respectively.
According to the design code of concrete cable-stayed bridges in China
[31], the threshold in Equation (3.4) is a = L/400 = 1.05 m. The evaluated
58 • BRIDGE RELIABILITY AND SERVICEABILITY
0.9
0.7 RADTT=0
RADTT=0.5%
RADTT=1
0.6
RADTT=1.5%
RADTT=2%
0.5
0 20 40 60 80 100
(a) Service period (year)
1
Maximum displacement (m)
0.9
0.8
0.7 RGVW=0
RGVW=0.5%
0.6 RGVW=1
0.5 RGVW=1.5%
RGVW=2%
0.4
0 20 40 60 80 100
(b) Service period (year)
10–6
RADTT
RGVW
Probability of failure
10–7
10–8
10–9
0 0.5 1 1.5 2
Annual linear growth rate (%)
10-6
RADTT=2%
RGVW=2%
Probability of failure
10-7
10-8
10-9
Unlimited 100% 75% 50% 25% 0
Threshold overload ratio
traffic growth ratio of 2 percent of RADTT and RGVW the probability of failure
decreased to 4.5 × 10−8 and 9.4 × 10−9, respectively.
The following interpretations can be made based on the numerical
results. First, the maximum traffic load effect is associated with the traffic
density and the GVW. Therefore, the intensive overloaded trucks in the
mid-span point of the bridge forms the critical loading scenario resulting in
the first-passage failure of the deformation of the cable-stayed bridge. The
growth rate of the ADTT increases the probability of exceedance due to the
intensive overloaded traffic flows, and then results in a higher probability
60 • BRIDGE RELIABILITY AND SERVICEABILITY
3.5 CONCLUSIONS
REFERENCES
[1] Deng, L., W. Wan, and Y. Yu. 2015. “State-of-the-Art Review on the
Causes and Mechanisms of Bridge Collapse.” Journal of Performance of
Constructed Facilities 30, no. 2, p. 04015005.
[2] Chen, S., and J. Wu. 2011. “Modeling Stochastic Live Load for Long-span
Bridge Based on Microscopic Traffic Flow Simulation.” Computers &
Structures 89, no. 9, pp. 813–24.
[3] Liu, Y., N.W. Lu, and X.F. Yin. 2016. “A Hybrid Method for Structural
System Reliability-Based Design Optimization and its Application to
Trusses.” Quality and Reliability Engineering International 32, no. 2,
pp. 595–608.
[4] Liu, Y., N.W. Lu, X.F. Yin, and M. Noori. 2016. “An Adaptive Support
Vector Regression Method for Structural System Reliability Assessment and
its Application to a Cable-Stayed Bridge.” P I Mech Eng O-J RIS 230, no.
2, pp. 204–19.
[5] Li, J., H. Hao, and J.V. Lo. 2015. “Structural Damage Identification with
Power Spectral Density Transmissibility: Numerical and Experimental
Studies.” Smart Structures and Systems 15, no. 1, pp. 15–40.
[6] Han, W., J. Wu, and C.S. Cai. 2015. “Characteristics and Dynamic Impact of
Overloaded Extra Heavy Trucks on Typical Highway Bridges.” Journal of
Bridge Engineering 20, no. 2, p. 05044011.
[7] Ministry of Communications and Transportation (MOCAT). 2004. General
Code for Design of Highway Bridges and Culverts. JTG D60-2004. Beijing,
China.
[8] Li, J., H. Hao, and Z. Chen. 2015. “Damage Identification and Optimal Sensor
Placement for Structures Under Unknown Traffic-Induced Vibrations.”
Journal of Aerospace Engineering 30, no. 2, p. B4015001.
[9] Marques, F., C. Moutinho, W.H. Hu, Á. Cunha, and E. Caetano. 2016.
“Weigh-in-motion Implementation in an Old Metallic Railway Bridge.”
Engineering Structures 123, pp. 15–29.
[10] O’Brien, E.J., and B. Enright. 2012. “Using Weigh-in-motion Data to
Determine Aggressiveness of Traffic for Bridge Loading.” Journal of Bridge
Engineering 18, pp. 232–39.
62 • BRIDGE RELIABILITY AND SERVICEABILITY
[11] Lu, N.W., M. Noori, and Y. Liu. 2016. “Fatigue Reliability Assessment
of Welded Steel Bridge Decks Under Stochastic Truck Loads via
Machine Learning.” Journal of Bridge Engineering. doi:10.1061/(ASCE)
BE.1943-5592.0000982
[12] Liu, Y., X.H. Xiao, N.W. Lu, and Y. Deng. 2016. “Fatigue Reliability
Assessment of Orthotropic Bridge Decks Under Stochastic Truck Loading.”
Shock and Vibration. doi:10.1155/2016/4712593
[13] Getachew, A., and E.J. Obrien. 2007. “Simplified Site-specific Traffic Load
Models for Bridge Assessment.” Structure and infrastructure Engineering 3,
no. 4, pp. 303–11.
[14] O’Brien, E.J., F. Schmidt, D. Hajializadeh, X. Zhou, B. Enright, C.C.
Caprani, S. Wilson, and E. Sheils. 2015. “A Review of Probabilistic Methods
of Assessment of Load Effects in Bridges.” Struct Saf 53, pp. 44–56.
[15] Enright, B., and E.J. O’Brien. 2013. “Monte Carlo Simulation of Extreme
Traffic Loading on Short and Medium Span Bridges.” Structure and
Infrastructure Engineering 9, no. 12, pp. 1267–82.
[16] Chen, X. 2014. “Estimation of Extreme Value Distribution of Crosswind
Response of Wind-Excited Flexible Structures Based on Extrapolation of
Crossing Rate.” Engineering Structures 60, pp. 177–88.
[17] AASTHTO. 2012. LRFD Bridge Design Specifications, 6th ed. Washington,
DC.
[18] O’Connor, A., and E.J. O’Brien. 2005. “Traffic Load Modelling and Factors
Influencing the Accuracy of Predicted Extremes.” Canadian Journal of Civil
Engineering 32, no. 1, pp. 270–78.
[19] Ruan, X., J. Zhou, X. Shi, and C.C. Caprani. 2016. “A Site-specific Traffic
Load Model for Long-span Multi-pylon Cable-stayed Bridges.” Structure
and Infrastructure Engineering. doi:10.1080/15732479.2016.1164724
[20] Song, J., and A. Der Kiureghian. 2006. “Joint First-passage Probability and
Reliability of Systems Under Stochastic Excitation.” Journal of Engineering
Mechanics 132, no. 1, pp. 65–77.
[21] Noori, M., M. Dimentberg, Z. Hou, R. Christodoulidou, and A. Alexandrou.
1995. “First-passage Study and Stationary Response Analysis of a BWB
Hysteresis Model Using Quasi-conservative Stochastic Averaging Method.”
Probabilistic Engineering Mechanics 10, no. 3, pp. 161–70.
[22] Chen, J.B., and J. Li. 2007. “The Extreme Value Distribution and Dynamic
Reliability Analysis of Nonlinear Structures with Uncertain Parameters.”
Structural Safety 29, no. 2, pp. 77–93.
[23] Khan, R.A., T.K. Datta, and S. Ahmad. 2005. “Reliability Analysis of Fan
Type Cable Stayed Bridges Against First Passage Failure Under Earthquake
Forces.” Journal of Seismology and Earthquake Engineering 7, no. 3,
pp. 147–57.
[24] Balafas, K., and A. Kiremidjian. 2015. “Reliability Assessment of the
Rotation Algorithm for Earthquake Damage Estimation.” Structure and
Infrastructure Engineering 11, no. 1, pp. 51–62.
FIRST-PASSAGE PROBABILITY OF THE DEFLECTION • 63
[25] Lydon, M., S.E. Taylor, D. Robinson, A. Mufti, and E.J. Brien. 2016.
“Recent Developments in Bridge Weigh in Motion (B-WIM).” Journal of
Civil Structural Health Monitoring 6, no. 1, pp. 69–81.
[26] Liu, Y., Y. Deng, and C.S. Cai. 2015. “Deflection Monitoring and Assessment
for a Suspension Bridge Using a Connected Pipe System: A Case Study in
China.” Structural Control and Health Monitoring 22, no. 12, pp. 1408–25.
[27] MOCAT. 2004. Limits of Dimensions, Axle Load and Masses for Road
Vehicles GB 1589-2004. Beijing, China: China Communications Press.
[28] Rice, S. 1944. “Mathematical Analysis of Random Noise.” The Bell System
Technical Journal 24, no. 1, pp. 46–156.
[29] Ditlevsen, O. 1994. “Traffic Loads on Large Bridges Modeled as White-
noise Fields.” Journal of Engineering Mechanics 120, no. 4, pp. 681–94.
[30] Cremona, C. 2001. “Optimal Extrapolation of Traffic Load Effects.”
Structural Safety 23, no. 1, pp. 31–46.
[31] MOCAT. 2007. Specifications for Design of Highway Cable-stayed Bridge
JTG/T D65-01. Beijing, China: China Communications Press.
CHAPTER 4
Dynamic Reliability
Evaluation of the
Serviceability of Cable-
Supported Bridges
Under Site-Specific
Heavy Traffic Loads
4.1 INTRODUCTION
4.2 TRAFFIC–BRIDGE INTERACTION
FORMULATION
where Mv, Cv, and Kv are the mass, the damping and the stiffness matrices
of the vehicle, respectively; Mb, Cb, and Kb are the mass, the damping and
the stiffness matrices of the bridge, respectively; uv and ub are the dis-
placement vectors of the vehicle and the bridge, respectively; Cbb, Cbv, Cvb,
Kbb, Kbv, Kvb, Fbr, and Fvr are the additional terms caused by the expansion
of the wheel-road contact force. All of these time-variant parameters will
change with the moving vehicle on the bridge.
The bridge behavior can be modeled linearly. That allows using
the modal superposition to reduce the number of DOFs of the system.
It is acknowledged that {Φ i }T [ M b ]{Φ i } = 1 , {Φ i }T [ K b ]{Φ i } = wi2 , and
[C b ] = 2wi hi [M b] , where ωi is the frequency of the ith mode shape of the
bridge, ηi is the damping ratio for the ith mode of the bridge. Therefore,
Equation (4.1) can be written in frequency domain:
D1 D2
Zv
qr1
(a)
Zv
qr2
Za1L Za3R
(b)
u b + Cb u b + K b u b = Feqwheel (4.3)
M b
nv nb
{ } (4.4)
Feqwheel (t ) = ∑ 1 − R j (t ) G j ⋅ ∑ hk [ x j (t ) + ak [ x j (t )d j (t )]]
j =1 k =1
where Feqwheel is the time-varying force vector on the bridge; Rj and Gj are the
dynamic load ratio and the gross vehicle weight of the jth vehicle; xj, and
dj are the longitudinal and transverse location of the gravity center of the
jth vehicle, respectively; hk and ak are the vertical and the torsional mode
shapes for the kth mode of the bridge model, respectively; nv and nb are the
number of vehicles and the number of adopted bridge modes, respectively.
It is worth observing that the road surface roughness condition is consid-
ered in the relative displacement. The effectiveness of the EDWL approach
has been demonstrated by Chen and Wu [4] with several applications.
It should be pointed out that the total number of vehicles changes with
time depending on the density of the stochastic traffic flow. Nevertheless,
it can be assumed to be stationary for dense traffic flow, and will be con-
sidered in the present study. With the EDWLs, the dynamic load effect of
the bridge under traffic loads can be efficiently calculated.
4.3 METHODOLOGY OF PROBABILISTIC
MODELING OF THE EXTREME
TRAFFIC LOAD EFFECTS
s′ ( x − m) 2
v( x) = exp − (4.5)
2ps 2s 2
where x is a random process, m and σ are the mean value and standard
deviation of x, respectively, and σ’ is the standard deviation of the deriv-
ative of x.
In practice, the v(x) can be fitted to histograms of number of cross-
ings in a unit time as shown in Figure 4.2(a, b). By dividing the entire
nonstationary process into several stationary interval processes, the
level-crossing rate of original process is supposed as a superposition of
the interval processes in proportions. Based on this assumption, the kernel
coefficients can be evaluated by a second-order polynomial function:
x
x
Down crossing
Interval length
(a) (b)
m p N ( x)
ln [v ( x) ] = ln ∑ i i = a0 + a1 x + a2 x (4.6)
2
i =1 T
where v ( x) is an estimated equivalent level-crossing rate of the nonsta-
m2 m 1
tionary process, a0 = ln(v0 ) − 2 , a1 = 2 , and a2 = are second-
2s s −2s 2
order polynomial coefficients that can be evaluated based on the histograms
of number of crossings.
For the purpose of extreme value extrapolation, the critical step of
the fitting is to find the optimal starting point and the number of inter-
vals. A starting point close to the tail is better for interpolation, while that
far from the tail is better for the extrapolation. A Kolmogorov–Smirnov
test suggested by Cremona [31] is an effective approach to optimize the
starting point. Based on the optimal starting point and class intervals, the
maximum value in a return period can be written as
where Rt is the return period, xmax is the maximum load effect correspond-
ing to the return period, mopt, σopt and v0,opt, represent the optimal mean
value, optimal standard derivation and optimal crossing rate, respectively.
In addition to the maximum value extrapolation based on the
level-crossing formula, first-passage failure probability evaluation is
another important measure for the assessment of bridge reliability under
traffic loads. First-passage failure is the best description of stochastic pro-
cess crossing the prescribed threshold during an interval time. Based on
the Rice’s level-crossing theory, the probability of failure, Pf , can be esti-
mated by the assumption of a Poisson distribution:
traffic load effect. In order to take into account the VBI induced traffic
load effect, a computational framework is presented integrating the VBI
and the level-crossing theory to extrapolate the maximum traffic load
effect of long-span bridges.
Flow chart of the computational framework is shown in Figure 4.3.
As depicted in Figure 4.3, the flow chart starts with the statistics of the
WIM measurements. The daily traffic flow can be simulated based on the
statistics of the WIM data. Subsequently, we can identify the critical traffic
loading scenario via the structural influence lines. This loading scenario
is used to compute the dynamic analysis of the vehicle–bridge coupled
vibration system. It is worth mentioning that the effect of vehicle spacing
in traffic density is a significant factor impacting the traffic load effects
on medium- to long-span bridges. The truck spacing instead of vehicle
WIM measurements
Statistics
Stochastic daily traffic loads
Structural influence lines
Modular procedures :
RRC EDWL
Convert truck weights into time-
variant moving loads
Transient analysis
Dynamic load effect histories
No
Enough number of days?
Yes
Level-crossing fitting to the
histograms of the number of crossings
Return period
Extrapolation of the maximum traffic
load effects
A WIM system utilizes scales or pressure sensors embedded into the road
pavement to measure physical properties of passing vehicles such as axle
weight and speed. Axles of travelling vehicles interrupt the magnetic signal
produced by the loop sensors, and therefore, the number of axles, the axle
spacing and the number of vehicles are recorded. In addition to the traffic
management, WIM can also provide a great amount of real traffic data
through loop sensors and piezo sensors. Site-specific WIM measurements
SERVICEABILITY OF BRIDGES UNDER TRAFFIC LOADING • 75
1
Empirical data
Lognormal
Probability density 0.8
0.6
0.4
m=6.21
0.2 s=1.40
0
0 1000 2000 3000 4000
(a) Vehicle spacing (m)
0.02
WIM data
GMM
0.015
Probability density
a1=0.24 a2=0.76
0.01 m1=390 m2=865
s1=74 s2=142
0.005
0
0 500 1000 1500 2000
(b) GVW (kN)
V1 V2 V3 V4 V5 V6
150
GVW (tonne)
100
50
0
0
200
400
600 4
800 3
2
Location (m) 1 Traffic lane
210
m
420
m
210
m
(a)
820
m
(b)
0.3
L/4
0.2 L/2
0.1
Deflection (m)
-0.1
-0.2
-0.3
-0.4
0 10 20 30 40 50 60
Time (min)
(a)
0.4
L/4
0.2 L/2
Vertical displacement (m)
-0.2
-0.4
-0.6
-0.8
-1
0 10 20 30 40 50 60
Time (min)
(b)
0.1
Cable-stayed bridge
Suspension bridge
0
-0.2
-0.3
0 L/4 L/2 3L/4 L
(a) Longitudinal location of the girder (m)
0.5
Cable-stayed bridge
Suspension bridge
0.4
0.3
RMS (m)
0.2
0.1
0
L/4 L/2 3L/4 L
(b) Longitudinal location of the girder (m)
0.1
Static
Dynamic
0
Deflection (m)
-0.1
Nd,1 = 6 Nd,2 = 4
-0.3
-0.4
0 21 42 63 84
Time (s)
(a)
0.5
Static
0.25 Dynamic
0
Deflection (m)
1st sample
Ns,1 = 1
-0.25
Nd,1 = 2 2nd sample
-0.5 Ns,2 = 0
Nd,2 = 2
-0.75
-1
0 20.5 41 61.5 82
(b) Time (s)
250
The cable-stayed bridge The suspension bridge
200
50
0
0.4 0.6 0.8 1 1.2 1.4
Deflection (m)
2
Cable-stayed bridge
Suspension bridge
1.7
Maximum deflection (m)
1.4
1.1
0.8
0.5
0 200 400 600 800 1000
Return period (years)
4.5.2 PARAMETRIC STUDIES
Since the traffic volume will grow with the development of global econ-
omy, the average daily truck traffic (ADTT) was selected herein for para-
metric studies. The ADTT is usually affected by numerous factors, and
is difficult to predict for the entire service period of a bridge. This study
SERVICEABILITY OF BRIDGES UNDER TRAFFIC LOADING • 83
10-4
Cable-stayed bridge
Suspension bridge
Probability of exceedance
10-5
10-6
10-7
10-8
0 20 40 60 80 100
Service period (years)
10-1
Cable-stayed bridge
Suspension bridge
10-2
Probability of exceedance
10-3
10-4
10-5
10-6
0 1% 2% 3% 4%
Annual linear growth rate of ADTT
10–1
Cable-stayed bridge
Suspension bridge
10–2
Probability of exceedance
10–3
10–4
10–5
10–6
Unlimited 100% 75% 50% 25%
Overload rate limit
ADTT. In addition, the growth rate of ADTT enlarges the gap between
the two values of the probability of failure. This demonstrates that a more
realistic traffic growth model is still important for the bridge reliability
assessment.
In addition to the traffic growth, traffic managements will also impact
the traffic flow. In general, an overload control measure will lead to a pos-
itive result. In order to study the influence of the overload control measure
on bridge serviceability, the threshold of overloading ratio is assumed to
be 25, 50, 75 and 100 percent. The thresholds of GVW for the 2-axle
and 6-axle trucks specified in the MOCAT (2004) are 200 kN and 550
kN, respectively. Based on the above assumption, the stochastic traffic
flow load model was updated with the defined overloading ratio, and
then the up-crossing rate and the probability of failure were re-evaluated.
Figure 4.14 shows the influence of the threshold of overloading ratio on
the probability of failure of the bridge.
As observed from Figure 4.14, the threshold of overloading ratio has
a remarkable influence on decreasing the structural probability of failure.
Even for a threshold of 100 percent, the probability of failure has a rapid
decrease in comparison with the unlimited state. In the case of gradual
traffic volume growth, the overload control measure is an effective way to
decrease serviceability risk of flexible cable-supported bridges subject to
long-term traffic loading.
SERVICEABILITY OF BRIDGES UNDER TRAFFIC LOADING • 85
4.6 CONCLUSIONS
The study presented in this chapter utilized a stochastic traffic flow load
model to estimate the reliability-based serviceability of cable-supported
bridges subject to long-term heavy traffic loading. The heavy traffic flow
was simulated based on site-specific weigh-in-motion measurements via a
cellular automation model. Subsequently, the deterministic load effect and
probabilistic modeling the extreme value were computed via the traffic–
bridge coupled vibration theory and via Rice’s level-crossing theory,
respectively. The case study of two typical cable-supported bridges has
demonstrated the effectiveness of the proposed computational framework.
The conclusions are summarized as follows:
challenges still remain to be addressed in future work. First, since the prob-
abilistic modeling of traffic load effects needs a large number of data, the
time-consuming traffic–bridge interaction analysis is still a bottleneck for
the computational efficiency. Second, utilizing the critical loading scenario
in the daily traffic flow, instead of the entire dense traffic flow, will save a
considerable amount of time. Third, the probability model and extrapolation
of the extreme traffic load effect should be compared with other approaches,
such as the generalized extreme value distribution and the peaks over value
approach. Finally, as aforementioned, a more realistic traffic growth model
should be considered instead of an idealized traffic growth model.
REFERENCES
[26] Chen, S.R., and C.S. Cai. 2007. “Equivalent Wheel Load Approach for
Slender Cable-stayed Bridge Fatigue Assessment Under Traffic and Wind:
Feasibility Study.” Journal of Bridge Engineering 12, no. 6, pp. 755–64.
[27] Deng, L., and C.S. Cai. 2010. “Development of Dynamic Impact Factor
for Performance Evaluation of Existing Multi-girder Concrete Bridges.”
Engineering Structures 32, no. 1, pp. 21–31.
[28] Deng, L., W. Wang, and Y. Yu. 2016. “State-of-the-art Review on the Causes
and Mechanisms of Bridge Collapse.” Journal of Performance of Constructed
Facilities 32, no. 2, 04045005. doi:10.1061/(ASCE)CF.1943-5509.0000731
[29] Rice, S. 1945. “Mathematical Analysis of Random Noise.” Bell System
Technical Journal 24, no. 1, pp. 46–156.
[30] Ditlevsen, O. 1994. “Traffic Loads on Large Bridges Modeled as White-
noise Fields.” Journal of Engineering Mechanics 120, no. 4, pp. 681–94.
[31] Cremona, C. 2001. “Optimal Extrapolation of Traffic Load Effects.”
Structural Safety 23, no. 1, pp. 31–46.
[32] Ministry of Communications and Transportation (MOCAT). 2015.
Specifications for Design of Highway Suspension Bridge JTG/T D65-05.
Beijing, China: China Communications Press.
[33] Lu, N., M. Noori, and Y. Liu. 2016. “Fatigue Reliability Assessment of
Welded Steel Bridge Decks Under Stochastic Truck Loads via Machine
Learning.” Journal of Bridge Engineering 22, no. 1, 04016105. 10.1061/
(ASCE)BE.1943-5592.0000982
[34] Liu Y., X. Xiao, N. Lu, and Y. Deng. 2016. “Fatigue Reliability Assessment
of Orthotropic Bridge Decks Under Stochastic Truck Loading.” Shock and
Vibration. doi:10.1155/2016/4712593
[35] Cai, C.S., J. Hu, S. Chen, and W. Zhang. 2015. “A Coupled Wind-vehicle-
bridge System and its Applications: A Review.” Wind and Structures 20,
no. 2, pp. 117–42.
[36] Yin, X., Y. Liu, and B. Kong. 2016. “Vibration Behaviors of a Damaged
Bridge Under Moving Vehicular Loads.” Structural Engineering and
Mechanics 58, no. 2, pp. 199–216.
CHAPTER 5
Lifetime Deflections of
Long-Span Bridges Under
Dynamic and Growing
Traffic Load
5.1 INTRODUCTION
5.2.1 TRAFFIC-BRIDGE INTERACTION
nv
na
{F (t )}
eq
wheel
{ }
= ∑ 1 − EDWL j (t ) / G j G j ⋅ ∑ hk [ x j (t ) + ak [ x j (t ) d j (t )]] (5.1)
j =1 k =1
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 93
Zv
¦È
r1
1 1 2 2
KuzL CuzL KuzL CuzL
Za1L Za2L
1 1 2 2
KlzL ClzL KlzL ClzL
x
(a)
Zv
¦È
r2
1 1 1 1
KuzL CuzL KuzR CuzR
Za1L Za2R
1 1 1 1
KlzL ClzL KlzR ClzR
y
(b)
where EDWLj (t) is the dynamic load of the j-th vehicle at time t; Gj, xj,
and dj are the self-weight, longitudinal location, and transversal location
of the center of gravity of the jth vehicle on the bridge, respectively; hk
and ak are the vertical and the torsional mode shapes for the kth mode of
the bridge, respectively; and nv and na are the number of vehicles on the
bridge and the number of axles of the j-th vehicle, respectively. The num-
ber of vehicles on the bridge changes with time depending on the stochas-
tic traffic flow. The effectiveness has been demonstrated by Chen and Wu
[34]. Note that the road-roughness coefficient (RRC) is considered in the
vertical displacement and velocity of the vehicle.
5.2.2 RICE’S FORMULA
s ( x − m) 2 1
v( x) = exp − = (5.2)
2 Às 2s 2 Rt
where x is the traffic load effect, m and σ are the mean value and stan-
dard deviation of the load effect, and s is the standard deviation of the
derivative of the load effects. In general, the level-crossing rate can be
expressed by a normalized rate indicated as the fitted histograms ver-
sus the summation of truncated remaining histograms. The critical step
of using Rice level crossing theory for extrapolating is to determine the
optimal starting point indicated as x0,opt and the optimal number of class
intervals indicated as Nopt [37]. In this regard, the conventional approach
is to utilize the Kolmogorov–Smirnov (K–S) statistics recommended and
described by Cremona [38] to check the confidence level of the predefined
starting point and number of class intervals. Eventually, the statistical
x
Up crossing
Level 1
t
Level 2
Down crossing
(a)
Optimal starting point
v(x)
Rice's fitting
Interval length
x
(b)
T
Fmax ( x, T ) = 1 − a = 1 − 1 − exp −
Rt
( x − m) 2
= exp −Tv0 exp − (5.3)
2s 2
8000
Actual traffic growth
4000
2000
E1 E2 E10
0
0 20 40 60 80 100
(a) Time (year)
E1 E2 ... En
P1= Fmax, 1 (x, Tint) P2= Fmax, 2 (x, Tint) Pn= Fmax, n (x, Tint)
(b)
be used for the load effect extrapolation. However, traffic volume could be
assumed stationary in a short period such as one or two years. This short
period is defined as an interval in this study.
An illustrative example of the interval traffic growth is shown in
Figure 5.3, where the curve is the volume of the average daily truck traf-
fics (ADTTs) accounting for the actual traffic growth, and the histograms
are the volumes of the ADTTs in the 10-year interval. It is obvious that
the traffic volume is constant in an interval period rather than growing
with the curve. The advantage of the interval growth model is that Rice’s
formula is appropriate to be used in each interval. The shortcoming of
the interval growth model is that the result is only an estimation and its
accuracy mostly depends on the number of intervals. Obviously, increase
of the intervals will lead to the improvement in the accuracy of the result,
but also leads to additional computational effects.
Rice’s formula as shown in Eqs. (4) and (5) are unavailable for traffic load
effects under growing traffic load because the traffic density is nonstation-
ary over the bridge lifetime. However, these formulas are effective in each
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 97
int ( x − mi ) 2
Nint N
G
where Fmax ( x, T ) is the maximum traffic load effect of a bridge in a refer-
ence period T considering traffic growth, Nint is the number of intervals, Tint
= T/Nint is the interval period, Fmax,i (x,T) is the CDF of the maximum traffic
load effects in the ith interval, and v0,i mi and σi are the mean crossing rate,
the mean value and the standard deviation of the load effects in the ith
interval, respectively. Thus, the maximum load effects in a return period
can be estimated based on the system CDF as shown in Equation (5.4) and
the general form of the CDF as shown in Equation (5.3), written as
−Tref
Rt = (5.5)
ln F
G
max ( x, Tref )
where z is a predefined threshold for the traffic load effect, and Fmax(z,Tref)
denotes the CDF of the maximum load effect in lifetime at the z value.
This equation provides a way for the quantification of the probability of
the traffic load effect exceeding a limit.
Based on the interval traffic growth model and the improved Rice’s for-
mula, a computational framework is presented for evaluating the maximum
load effects of long-span bridges using WIM measurements. The flowchart
of the computational framework is shown in Figure 5.4. The entire pro-
cedure outlined in the flowchart is mainly composed of three categories
including the traffic load simulation, the load effect computation, and the
probabilistic extrapolation. Illustrations of the procedures in the categories
are elaborated in the following.
The first module as depicted in Figure 5.4 is the traffic load simulation
based on recorded WIM data. With available WIM data, filtering proce-
dures should be conducted to exclude the invalid data and select the effec-
tive data that contribute to the maximum traffic load effect. In the present
study, lightweight cars were removed from the recorded data, since most
of the critical loading scenarios are usually composed of dense traffic
flows with a high proportion of heavy trucks. The statistical parameters
of the traffic flow can be divided into two groups [17]: (a) those modeling
the individual vehicle feature in small-scale (i.e., the vehicle configura-
tion, the gross vehicle weight, the vehicle spacing and the driving speed);
and (b) those modeling the traffic feature in large-scale (i.e., proportions
of vehicle types and traffic volume). The most critical parameter is the
vehicle spacing defining the vehicle gap, that is, the space between two
vehicles in the same driving lane. The vehicle spacing is a unique and
critical factor for the traffic flow simulation on long-span bridges, and is
usually measured by the headway that is time variant depending on the
traffic density [19]. The extreme traffic load effects are mostly induced
by dense traffic flows with small vehicle spacings. Therefore, the PDF of
the vehicle spacing in dense traffic flows is one of the important factors
impacting the maximum traffic loading on a long-span bridge.
In general, traffic loads on a bridge can be simulated with a math-
ematical model in time domain or in space domain. A stochastic traffic
flow is one of these mathematical models composed of individual vehicles
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 99
WIM measurements
Filtering
Statistics of traffic flows
Stochastic traffic simulation
Daily critical traffic-loading scenarios
Probabilistic extrapolation:
Counting the
number of crossings
Fitting the level-crossing rates to the
histograms of the number of level-crossings
Return period
Extrapolation of the maximum traffic load
effects
Threshold value
Lifetime probability of exceedance of
the limit
influence line analysis aimed at finding out the maximum load effects in
the daily stochastic traffic flow. Based on this assumption, a step by step
search strategy is adopted to identify the critical loading scenario. These
procedures are: (1) generating the stochastic traffic flows via MCS and the
statistics of the WIM data; (2) specifying the effective range of the loading
scenario on the bridge according to the bridge length; (5.3) moving the
predefined range forward along the simulated daily stochastic traffic flows
to calculate the static traffic load effect; (5.4) identifying the maximum
load effect and the corresponding loading scenario; and (5.5) repeating
steps (5.3) to (5.4) for the remaining daily traffic flows.
The second module is the traffic load effect computation. There are
two factors considered in this module including the VBI and the traffic
growth. For VBI, the dynamic vehicle load of the bridge under the critical
loading scenario can be evaluated using the EDWL approach with the
consideration of the RRC. The result is a time history that will be used
for counting the number of level crossings. For the traffic growth, the
interval traffic growth model is used to simulate traffic growth with an
AGR. The daily maxima in each interval are estimated using the static
influence lines.
The third module is the probabilistic extrapolation. The first step is
to count the number of level crossings based on the estimated traffic load
history or the interval daily maxima. The level-crossing rate as shown in
Figure 5.2(b) can be fitted to the histograms of the number of crossings.
Subsequently, the maximum traffic load effect over a return period can be
extrapolated based on Rice formula in Equation (5.3) or the interval model
in Equation (5.4). Finally, the probability of exceedance of the predefined
limit can be evaluated from the CDF of the maximum value.
Obviously, there are some key points in the proposed method. Firstly,
more number of traffic intervals will lead to a more accurate extrapola-
tion, but also leads to more computations. Subsequently, since the actual
traffic volume in each traffic interval will grow instead of being constant,
which is assumed in the proposed approach, the extrapolated value will be
slightly under-conservative. In addition, consideration of lightweight cars
should be further developed.
5.5 VERIFICATION EXAMPLES
80
1st interval
5th interval
Annual number of crossings
60 10th interval
40
20
0
60 65 70 75 80
(a) GVW (t)
15
Exact values
1000-year return period
No-growth ACGR=4.1%
10
GEV GEV
Rice Rice
–log[–log(F)]
–5
60 65 70 75 80
(b) GVW (t)
GVW=50 t Gi ~N(50, 5) t
50 mm
Deflection: y = Gisin(x¦Ð/L)
50
1st interval
5th interval
40
10th interval
Annual number of crossings
30
20
10
0
50 60 70 80 90 100 110
Deflection (mm)
15
1000-year return period
GEV (full)
10
Rice (full)
–log[–log(Fmax)]
GEV (interval)
No=growth
5 Rice (interval)
AGR=1%
0
–5
50 70 90 110 130
Deflection (mm)
data rather than the lower data. This maybe due to the insufficient number
of intervals as shown in Figure 5.3(b) in capturing a higher probability
of exceedance.
5.6 CASE STUDY
A suspension bridge and its WIM data are utilized to demonstrate the
effectiveness of the proposed computational framework. Dynamic char-
acteristics and the traffic growth are considered to investigate their influ-
ences on the probabilistic extrapolation.
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 105
5.6.1 WEIGH-IN-MOTION MEASUREMENTS OF A
SUSPENSION BRIDGE
200
150
100
50
0
0:00 4:00 8:00 12:00 16:00 20:00 24:00
(a) Time
0.1
WIM data
Lognormal distribution
0.08
Probability density
0.06
0.04
0
0 1000 2000 3000 4000
(b) Truck spacing (m)
makes sense in the context of probabilistic domain, since the PDF of the
truck spacing was fitted to the actual WIM data. Even though the truck
spacing does not represent an actual traffic state, the PDF has captured the
statistical characteristics of the trucks in the actual traffic flows.
As the first task, a commercial program ANSYS was utilized to build the
bridge finite element (FE) model as shown in Figure 5.9. In the FE model,
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 107
Table 5.2. The first five order mode frequencies of the suspension bridge
Mode frequency
FE Monitored Error
Order model data (%) Illustration
1 0.131 – – 1st asymmetric transversal bending
2 0.1781 0.1849 −3.68 1st asymmetric vertical bending
3 0.2208 0.2492 −11.40 1st symmetric vertical bending
4 0.3106 0.3049 1.87 2nd asymmetric vertical bending
5 0.4074 0.3975 2.49 2nd symmetric vertical bending
WIM system
89 m
L=
82
0m 145 m
the steel stiffening steel-box girders and the concrete pylons were modeled
by beam elements, the hangers and the cables were modeled by link ele-
ments, the bridge tower elements and the cable elements were separated
and the top joints were connected by a set of coupled degrees of free-
dom. It is acknowledged that long-span bridges are geometrically nonlin-
ear, especially under heavy traffic loads. The present study, however, is
limited to a linear analysis as a first step for implementing the improved
stochastic analysis. The advancement of the improved stochastic analysis
can be extended to the nonlinear case, where the computational efficiency
should be further developed. Table 5.2 summarizes the first 5 fundamental
mode characteristics of the FE model. The vehicle physical properties in
this case study were adopted from Yin et al. [42]. The road surface rough-
ness conditions (RRCs) in this study were defined by the International
Organization for Standardization (1995). The RRCs for classifications
of “Good,” “Average,” and “Poor” are 32×10−6, 128×10−6, and 512×10−6,
respectively. RRC coefficients were simulated by inverse Fourier transfor-
mation approach [3] in time domain.
108 • BRIDGE RELIABILITY AND SERVICEABILITY
0.02
Static
0.01 Dynamic
Vertical deflection (m)
0
-0.01
-0.02
L/2 point
-0.03
3L/8 point
L/4 point
-0.04
0 L/4 L/2 3L/4 L
Vehicle location on the bridge (m)
As the second task, the dynamic traffic load effects at the critical points
of the bridge were simulated. Initially, a 2-axle truck with GVW = 10 t, a
driving speed v = 20 m/s, and the RRC = “Good” were considered to show
the static and dynamic deflections of the bridge girders. Figure 5.10 plots
the vertical deflections of the three potential points versus truck loading
position on the bridge, where the bridge self-weight effects were excluded.
It is observed that the most critical point is the L/4 (quarter-span) point.
The maximum deflection due to static and dynamic loads are 0.029 and
0.032 m, respectively. Therefore, the subsequent investigation focuses on
the quarter-point of the bridge.
Subsequently, the critical traffic loading scenarios were identified
from the simulated daily stochastic traffic flows. An illustration exam-
ple for generating the critical loading scenario on a 4-lane bidirectional
bridge is shown in Figure 5.11, where traffic lanes of 1 and 2 are in the
same direction, while the traffic lanes of 3 and 4 are in the opposite direc-
tion. A daily maximum deflection was identified from the static deflection
histories in Figure 5.11(a), and the corresponding critical traffic loading
scenario was extracted from the daily stochastic traffic load as shown in
Figure 5.11(b). Obviously, the critical loading scenario is a small part
(0.1%) of the daily traffic flow. As a result, the time-consuming com-
putation due to the step-by-step integration will be greatly reduced by
utilizing the critical traffic loading scenario rather than the entire daily
traffic stochastic flows.
A total number of 1,000-day traffic flows were simulated via
Monte-Carlo simulation for the current traffic condition. Each critical traf-
fic loading scenario was identified according to the maximum total GVW
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 109
0.6
0.4
0.2
Deflection (m)
0
-0.2
-0.4
-0.6
Daily maxima
-0.8
0 2 4 6 8 10
(a) Time (h)
120
80
GVW (t)
V1
40 V2
V3
0 V4
0 V5
205 V6
410
615 4
3
820 2
(b) Location (m) 1 Traffic lane
on the bridge. Suppose these critical traffics are free flowing with a constant
velocity v = 20 m/s. Subsequently, the traffic-bridge coupled vibration anal-
ysis was conducted by utilizing the EDWL approach. Figure 5.12 shows
the difference between the number of crossings of static and dynamic his-
tories, where the numbers of down crossings at the deflection level a =
−0.667 m was counted, and Nistatic and Nidynamic are the number of crossings
for the static and the dynamic histories, respectively. The deflection level
is different from the deflection threshold. The deflection level was utilized
to count the number of crossings, but the deflection threshold was utilized
to evaluate the probability of exceedance. It is observed, from Figure 5.12,
that the dynamic histories fluctuate around the static histories. This makes
the numbers of level crossings different. The number of the dynamic cross-
ings is always higher than that of the static crossings. This numerical result
shows the advantage of Rice’s formula in capturing the dynamic effects.
110 • BRIDGE RELIABILITY AND SERVICEABILITY
0.4
Static
0.2 Dynamic
0
v = 20 m/s
Deflection (m)
-0.4
Dynamic fitting
x0 = 0.584
v0 = 1001.9
1000
µ = 0.4889
σ = 0.1844 x = 0.584
0
500 v0 = 701
µ = 0.4813
σ = 0.1825
0
0 0.2 0.4 0.6 0.8 1 1.2
Deflection level (m)
1.5
1.3
Static
1.2 Good
Average
Poor
1.1
0 200 400 600 800 1000
Return period (years)
AGR=1% AGR=2%
300
200
100
0
100
80
60
40
20 1.5
0.9 1.2
0 0.3 0.6
Time (years) 0
Deflection level (m)
1.9
AGR=0
AGR=1%
Extrapolated deflection (m)
1.8 AGR=2%
AGR=3%
1.7
1.6
1.5
1.4
0 20 40 60 80 100
Service period (years)
8
AGR = 0
6 AGR = 1%
–log[–log(–Fmax(x,100)]
AGR = 2%
4 AGR = 3%
2 Rt =1000 years
0
-2
-4
1 1.2 1.4 1.6 1.8 2
(a) Deflection level (m)
100
L/400 = 2.05 m
10–2
Probability of exceedance
10–4
10–6
AGR=0
–8 AGR=1%
10
AGR=2%
AGR=3%
10–10
1.2 1.4 1.6 1.8 2 2.2
(b) Deflection level (m)
5.7 CONCLUSIONS
REFERENCES
[1] Fu, G., and J. You. 2009. “Truck Loads and Bridge Capacity Evaluation
in China.” Journal of Bridge Engineering, 327–35. 10.1061/(ASCE)
BE.1943-5592.0000006
116 • BRIDGE RELIABILITY AND SERVICEABILITY
[2] Zhu, S., D. Levinson, H. Liu, and K. Harder. 2010. “The Traffic and
Behavioral Effects of the I-35W Mississippi River Bridge Collapse.”
Transportation Research Part A: Policy and Practice 44, no. 10, pp. 771–84.
[3] Deng, L., and C.S. Cai. 2009. “Bridge Scour: Prediction, Modeling,
Monitoring, and Countermeasures—Review.” Practice periodical on
structural design and construction 15, no. 2, 125–34. 10.1061/(ASCE)
SC.1943-5576.0000041.
[4] Han, W., J. Wu, C.S. Cai, and S. Chen. 2014. “Characteristics and Dynamic
Impact of Overloaded Extra Heavy Trucks on Typical Highway Bridges.”
Journal of Bridge Engineering 20, no. 2, 05014011. 10.1061/(ASCE)
BE.1943-5592.0000666.
[5] Deng, L., W. Wang, and Y. Yu. 2016. “State-of-the-art Review on the
Causes and Mechanisms of Bridge Collapse.” Journal of Performance of
Constructed Facilities 30, no. 2. 10.1061/(ASCE)CF.1943-5509.0000731
[6] European Commission. Directorate General for Energy and Transport. 2008.
European Energy and Transport. Trends for 2030. Update 2007.
[7] Zhou, Y., and S. Chen. 2014. “Dynamic Simulation of a Long-span Bridge-
traffic System Subjected to Combined Service and Extreme Loads.”
Journal of Structural Engineering 141, no. 9, 04014215. 10.1061/(ASCE)
ST.1943-541X.0001188.
[8] Frangopol, D., A. Strauss, and S. Kim. 2008. “Bridge Reliability Assessment
Based on Monitoring.” Journal of Bridge Engineering 13, no. 3, 258–70.
10.1061/(ASCE)1084-0702.
[9] Li, S., S. Zhu, Y. Xu, Z. Chen, and H. Li. 2012. “Long-term Condition
Assessment of Suspenders Under Traffic Loads Based on Structural
Monitoring System: Application to the Tsing Ma Bridge.” Structural Control
and Health Monitoring 19, no. 1, pp. 82–101.
[10] Xia, M., C.S. Cai, F. Pan, and Y. Yu. 2016. “Estimation of Extreme Structural
Response Distributions for Mean Recurrence Intervals Based on Short-term
Monitoring.” Engineering Structures 126, pp. 121–32.
[11] Nowak, A.S. 1995. “Calibration of LRFD Bridge Code.” Journal of
Structural Engineering 121, no. 8, 1245–51. 10.1061/(ASCE)0733-9445.
[12] Kwon, O.S., E. Kim, and S. Orton. 2010. “Calibration of Live-load Factor
in LRFD Bridge Design Specifications Based on State-specific Traffic
Environments.” Journal of Bridge Engineering, 812–19. 10.1061/(ASCE)
BE.1943-5592.0000209.
[13] Caprani, C.C. 2012. “Calibration of a Congestion Load Model for
Highway Bridges Using Traffic Microsimulation.” Structural Engineering
International 22, no. 3, pp. 342–48.
[14] Enright, B., and E.J. OBrien. 2013. “Monte Carlo Simulation of Extreme
Traffic Loading on Short and Medium Span Bridges.” Structure and
Infrastructure Engineering 9, no. 12, pp. 1267–82.
[15] Guo, T., D.M. Frangopol, and Y. Chen. 2012. “Fatigue Reliability Assessment
of Steel Bridge Details Integrating Weigh-in-motion Data and Probabilistic
Finite-element Analysis.” Computers & Structures 112, pp. 245–57.
LIFETIME DEFLECTIONS OF LONG-SPAN BRIDGES • 117
System Reliability
Evaluation of in-Service
Cable-Stayed Bridges
Subject to Cable
Degradation Via a Machine
Learning Based Tool
6.1 INTRODUCTION
Both parallel wires and strands are conventional types of stay cables in
bridge engineering. A parallel wire cable consists of straight parallel round
wires inside a polystyrene pipe. In addition to the material characteristics,
the strength of parallel wires is also associated with the length of wires,
the number of wires, and the cable degradation [21]. Initially, this study
conducts mathematical formulation of cable degradation due to the effects
of cable length and the number of wires.
In general, a parallel wire cable can be modeled in a parallel–series
system as shown in Figure 6.1, where each wire can be simulated as
a series system and the wires work together as a parallel system. For the
series system, the individual wires of a stay cable can be simulated with
correlative elements depending on the length of the wire. The material
properties and defects in the wires are considered by a correlation length
defined as L0 in the series system. The cable strength decreases result-
ing from a shorter correlation length or a longer wire length. Therefore,
both the correlation length and the wire length should be considered in the
cable strength model. Distribution function of the strength of a wire can be
written by a Weibull distribution function
zk
FZ ( z ) = 1 − exp − l (6.1)
u
In
pa
ral
0m
el l
10
L=
.5m
=0
L0
s
erie
Ins
by a function of the scale factor, λ, which is the ratio between the length
of the wire specimen and the correlation length. An experimental study of
a 100 m long wire conducted by Faber et al. [22] indicated that the mean
values of the wire strength for undamaged cable (λ = 3) and the corroded
wire (λ = 200) were 1748 and 1650 MPa, respectively. The variability of
the wire strength is negligible according to Faber’s conclusion.
In addition to the series system of an individual wire, a stay cable con-
sists of numerous wires in parallel as shown in Figure 6.1. In the parallel
system, the increase of the number of wires in a cable leads to a reduction
of the mean strength of each wire, which is the so-called Daniel’ effect.
The general reduction can be up to about 8%, and the deviation of the
Daniel’s effect is negligible.
Since high stress cables are prone to be corroded [23], the stay cables became
rusty and fractured under long-term influence of corrosion and cyclic
stresses. Thus, cable degradation is a common phenomenon in existing
cable-supported bridges. Deterioration of cable wires takes different forms
including stress corrosion cracking, pitting, corrosion fatigue and hydrogen
embrittlement, which compromise the strength and ductility of wires lead-
ing to a reduced service life of bridge cables [24]. This study considers cable
degradation resulting from atmospheric corrosion and fatigue damage.
In the concept of fatigue damage accumulation, the failure times of
the undamaged wires under a mean stress range can be assumed to be
identical and independent. The wire with the smallest failure times breaks
SYSTEM RELIABILITY OF CABLE-STAYED BRIDGES • 123
firstly in the system. Subsequently, the remaining wires have stress redis-
tributions. Therefore, the probability distribution function of the initial
failure times is written as [25]
∆S a a
eq N m
FN ( Seq , N ) = 1 − exp (6.2)
rc K
where ΔSeq and N are the equivalent stress range and the number of stress
cycles, respectively; α, m, and K are unknown coefficients that can be
estimated by experimental tests, rc is a parameter associated with the
cross-sectional area of the cable. Faber et al. [26] utilized the ultrasonic
inspection technology to investigate the fatigue stress degradation, and
then provided the degradation functions for the main cables, hanger
cables, and stay cables for non-corroded and corroded wires. Based on
Faber’s results, the strength coefficients of non-corroded and corroded
wires were interpolated for 20-year service period as shown in Figure 6.2.
In Figure 6.2, the non-corroded cable is associated with only fatigue
damage, and the corroded cable is associated with both fatigue and cor-
rosion. It is observed that the strength coefficients of the stay cable in the
20-year service period resulting from fatigue and fatigue-corrosion effects
are 0.928 and 0.751, respectively. In addition, the curve for the fatigue
effect is close to linear, while the curve accounting for fatigue-corrosion
effect is nonlinear. Therefore, it is obvious that corrosion results in a rapid
decrease of the cable strength.
0.95
Strength coefficient
0.9
0.75 Non-corroded
Corroded
0.7
0 5 10 15 20
Service period (years) s
2
M P A2
= 1− (6.3)
MP PP 4 wZ x
cable. For the ductile failure, such as the bending failure of the girders or
towers, the subsystem is generated by adding a plastic hinge at the failure
location. One of the search techniques for the potential failure components
is the branch-and-bound method that is the so-called β-unzipping method
[29]. Suppose a system consists of n components, and k-1 components
denoted as r1, r2,…, rk-1 failed. The conditional reliability index of the kth
component is written as:
( )
br(kk/) = br(kk/)r1, r2,rk −1 = Φ −1 P Er(kk/) (6.4)
where Er(kk/) is the event of the kth component failure at the kth failure
phase, P() denotes the probability of the event, Φ −1 () is an inverse cumu-
lative distribution function, and br(kk/)r1, r2,rk −1 is a conditional reliability
index of the kth potential component at the kth failure phase. The condi-
tion to be the potential failure component is [30]
β r(kk/) = β min
(k )
+ ∆β (6.5)
where bmin (k )
is the minimum reliability index at the kth failure phase,
and Δβ equals to 3 and 1 at the first and the following searching phases,
respectively.
As observed from the previous reasoning, the beam–column effect is
considered in Equation (6.4), and the failure sequence search criteria is
shown in Equations (6.4 and 6.5). In addition to these unique characteris-
tics illustrated earlier, the cable degradation will add to the time-consuming
computation of the system reliability. Therefore, special attention should
be paid for utilizing an efficient computational framework.
Component reliability
evaluations
Updating the finite-
Cable degradation
element model
model as shown in
Removing the potential failed
Fig. 2
components separately
No
Is the new structure failed?
Yes
System reliability evaluation
based on the parallel-series
model
No, then
T=T+1
T>Ts ?
Yes
Lifetime sytem reliability
cable strength has changed, because the cable degradation may change the
potentially failed components that form the failure sequence. Finally, if the
service period, Ts, is reached, the entire procedure will stop and the life-
time system reliability indices are generated as output. The cable degrada-
tion induced system reliability decreases are reflected by the time-variant
system reliability indices.
The critical step depicted in Figure 6.3 is to return to the sampling
training samples after the cable strength is updated. This procedure sug-
gests that it is inappropriate to only update the reliability index of indi-
vidual cables without reevaluating the failure sequences. Instead, the
seemingly additional computational effort is essential to capture the main
failure sequences. Such deduction will be demonstrated in a case study.
CSRA
6.5 CASE STUDIES
Rupture
15m×2
C1 T1 C4
C2 C3 Plastic change
T2
G1 G2 G3 G4 G5 G6 G7 G8 G9 G10 G11
25m
15m×12
Figure 6.6. Event trees of the cable-stayed bridge: (a) without cable degrada-
tion; (b) with a cable degradation coefficient of 20%.
SYSTEM RELIABILITY OF CABLE-STAYED BRIDGES • 131
Figure 6.6. (Continued)
3.5
2.5 Fatigue
Fatigue and Corrosion
2
0 5 10 15 20
Service period (years)
Figure 6.7. System reliability index of the cable-stayed bridge subject
to cable degradation.
Degration
210 m 420/2 m
7m 32×6 m 32×6 m+1 m 30.3 m
141
77 m
1 34
1
Cs34 Cs P1 Cm Cm
P3 P1
40 m 37 m
P2
34
Gs
1
Gs Gm
1 34
Gm
Symmetric P2
(a) (b)
fourth longest cables are shown in Table 6.1, where Ns is the number of
wires in a stay cable, As is the cross sectional area of the stay cable, Es,Ernst
is the equivalent modulus of elasticity of the stay cable, P0 is the cable
force, and Pb is the strength of the cable. More physical properties of the
parallel steel wires can be found in MOCAT [35]. The random variables
include the modulus of elasticities, equivalent densities, cross-sectional
areas and the strength of the girders, pylons and cables. More details on
the stochastics of the bridge can be found in Liu et al. [30].
A finite element model shown in Figure 6.9 was built based on the com-
mercial program ANSYS, where the cables were simulated by LINK180
elements, and the girders and pylons were simulated by BEAM188 ele-
ments. The traffic loads were considered as uniformly distributed forces
in the mid-span. The critical failure components are the longest cables
according to Liu et al. [30]. Taking the Cm31, Cm32, and Cm33 cables and the
P2 and P3 girders as examples, their dynamic response under the sudden
failure of the Cm34 cable were analyzed. The response histories are shown in
Figure 6.10. As observed from Figure 6.10, both the cable forces and the
bending moments increase and fluctuate under the case of sudden failure
of the Cm34 cable. The dynamic amplification factor for the maximum cable
force and the maximum bending moment of the girder are 1.03 and 1.09,
respectively. In addition, the components close to the failure cables have
stronger responses.
With the deterministic analysis results from the finite-element
model, the response surface function can be approximated by the learn-
ing machines. In this case, the implicit performance functions were for-
mulated by the SVR model. In addition, the SVR model can be updated
by removing the failed components and reevaluating the response of the
new structure. Response function of the Cm33 cable force associated with
the live load, the cross-sectional area of cables and the Cm34 cable rupture
are shown in Figure 6.11, where FC is the cable force of Cm33 cable, Q is
33
m
the uniformly distributed forces of the live load in the mid-span, and As is
the cross-sectional area as shown in Table. 6.1.
Based on the proposed framework, the event trees of the bridge were
established accounting for the fatigue and corrosion effects shown in Figure
6.1. Since there are 34 pairs of cables for each side of a pylon and they have
high correlation factors, only the longest cables including Cm34 and Cs34 were
selected for the first layer of the event tree to simplify the computation. Two
event trees of the bridge in different service periods are shown in Figure 6.12.
As depicted in Figure 6.12, there are two dominant failure sequences.
The first failure sequence is the strength failure of the mid-span cables (Cm34
and Cm33) followed by the bending failure of the mid-span girder (Hinge @
P3). The second failure sequence is the strength failure of the side-span
SYSTEM RELIABILITY OF CABLE-STAYED BRIDGES • 135
7000
33
C
m
6500 32
C
m
5500
5000
0 5 10 15 20 25
Time (s)
(a)
-2
-3 P
3
Bending moment (MN.m)
-4
-5 P
2
-6
-7
0 5 10 15 20 25
Time (s)
(b)
34
After C m ruptures
6800
6700
F 33 (kN)
6600
m
C
6500
34
Before C ruptures
6400 m
80
12
70
11
60 10
50 7
Q (kN.m-1) 8 As (×10-3 m2)
Figure 6.11. Response surfaces of the Cm33 cable force due to Cm34 rupture.
136 • BRIDGE RELIABILITY AND SERVICEABILITY
cables (Cs34 and Cs33) followed by the bending failure of the pylons (Hinge
@ P2). Accounting for the 6 failure sequences and the symmetry of the
bridge, the system reliability indices of the bridge are 7.60 and 6.41 for the
T = 0 and T = 20 years. The system reliability index of the bridge in the 40
years of service period is shown in Figure 6.13.
SYSTEM RELIABILITY OF CABLE-STAYED BRIDGES • 137
4 Fatigue
Fatigue and corrosion
3
0 10 20 30 40
Service period (years)
6.6 CONCLUSIONS
cables due to the cable corrosion. As a result, the structural system reli-
ability experiences a significant decrease in the period that the cable reli-
ability is inferior to the one of the critical girder. For the long-span bridge,
the system reliability index decreases to the threshold value in the 29-year
service period taking into account both fatigue and corrosion. Since the
corrosion leads to the rapid decrease of the cable strength as well as the
structural system reliability, implementing effective anti-corrosion mea-
sures and replacing deeply corroded cables are necessary for ensuring life-
time system safety of cable-stayed bridges.
There still remain some challenges in the reliability assessment of
long span bridges that require further study. First, site-specific measure-
ments of the cable strength need to be utilized for establishing a more
reasonable cable degradation model. Second, a more complete event tree
of the cable-stayed bridge instead of taking the longest cables as the first
order, as adopted in the present study, should be formulated. Finally, cor-
relation coefficients need to be considered to make the estimation of the
system reliability more accurate.
REFERENCES
[1] Yang, O., H. Li, J. Ou, and Q.S. Li. 2013. “Failure Patterns and Ultimate
Load-carrying Capacity Evolution of a Prestressed Concrete Cable-stayed
Bridge: Case Study.” Advances in Structural Engineering 16, no. 7,
pp. 1283–96.
[2] Li, H., C.M. Lan, Y. Ju, and D.S. Li. 2011. “Experimental and Numerical
Study of the Fatigue Properties of Corroded Parallel Wire Cables.” Journal
of Bridge Engineering 17, no. 2, pp. 211–20.
[3] Mehrabi, A.B., C.A. Ligozio, A.T. Ciolko, and S.T. Wyatt. 2010. “Evaluation,
Rehabilitation Planning, and Stay-cable Replacement Design for the Hale
Boggs Bridge in Luling, Louisiana.” Journal of Bridge Engineering 15,
no. 4, 364–72. 10.1061/(ASCE)BE.1943-5592.0000061
[4] Marjanishvili, S.M. 2004. “Progressive Analysis Procedure for Progressive
Collapse.” Journal of Performance of Constructed Facilities 18, no. 2, pp. 79–85.
[5] Cao, Y., G.W. Vermaas, R. Betti, A.C. West, and P.F Duby. 2003. “Corrosion and
Degradation of High-strength Steel Bridge Wire.” Corrosion 59, no. 6, pp. 547–54.
[6] Wolff, M., and U. Starossek. 2009. “Cable Loss and Progressive Collapse in
Cable-stayed Bridges.” Bridge Structures 5, no. 1, pp. 7–28.
[7] Xu, J., and W. Chen. 2013. “Behavior of Wires in Parallel Wire Stayed Cable
Under General Corrosion Effects.” Journal of Constructional Steel Research
85, pp. 40–47.
SYSTEM RELIABILITY OF CABLE-STAYED BRIDGES • 139
[8] Mozos, C.M., and A.C. Aparicio. 2010. “Parametric Study on the Dynamic
Response of Cable Stayed Bridges to the Sudden Failure of a Stay, Part I:
Bending Moment Acting on the Deck.” Engineering Structures 32, no. 10,
pp. 3288–300.
[9] Mozos, C.M., and A.C. Aparicio. 2011. “Numerical and Experimental Study
on the Interaction Cable Structure During the Failure of a Stay in a Cable
Stayed Bridge.” Engineering Structures 33, no. 8, pp. 2330–41.
[10] Zhou, Y., and S. Chen. 2014. “Time-progressive Dynamic Assessment of
Abrupt Cable-breakage Events on Cable-stayed Bridges.” Journal of Bridge
Engineering 19, no. 2. 159–71. 10.1061/(ASCE)BE.1943-5592.0000517
[11] Zhou, Y., and S. Chen. 2015. “Numerical Investigation of Cable Breakage
Events on Long-span Cable-stayed Bridges Under Stochastic Traffic and
Wind.” Engineering Structures 105, pp. 299–315.
[12] Aoki, Y., H. Valipour, B. Samali, and A. Saleh. 2013. “A Study on Potential
Progressive Collapse Responses of Cable-stayed Bridges.” Advances in
Structural Engineering 16, no. 4, pp. 689–706.
[13] Cheng, J., and R.C. Xiao. 2005. “Serviceability Reliability Analysis of Cable-
stayed Bridges.” Structural Engineering and Mechanics 20, no. 6, pp. 609–30.
[14] Li, H., S. Li, J. Ou, and H. Li. 2012. “Reliability Assessment of Cable-stayed
Bridges Based on Structural Health Monitoring Techniques.” Structure and
Infrastructure Engineering 8, no. 9, pp. 829–45.
[15] Bruneau, M. 1992. “Evaluation of System-reliability Methods for
Cable-stayed Bridge Design.” Journal of Structural Engineering 118, no. 4,
1106–20. 10.1061/(ASCE)0733-9445
[16] Estes, A.C., and D.M. Frangopol. 2001. “Bridge Lifetime System Reliability
Under Multiple Limit States.” Journal of Bridge Engineering 6, no. 6,
523–28. 10.1061/(ASCE)1084-0702
[17] Liu, Y., N. Lu, N. Yin, and M. Noori. 2016a. “An Adaptive Support Vector
Regression Method for Structural System Reliability Assessment and its
Application to a Cable-stayed Bridge.” P. I. Mech O: J. Ris., 230, no. 2,
pp. 204–19.
[18] Thoft-Christensen, P., and Y. Murotsu. 1986. Application of Structural
Systems Reliability Theory. Heidelberg, Germany: Speinger-Verlag Berlin.
[19] Lee, Y.J., and J. Song. 2011. “Risk Analysis of Fatigue-induced Sequential
Failures by Branch-and-bound Method Employing System Reliability
Bounds.” Journal of Engineering Mechanics 137, no. 12, 807–21. 10.1061/
(ASCE)EM.1943-7889.0000286
[20] Kim, D.S., S.Y. Ok, J. Song, and H.M. Koh. 2013. “System Reliability
Analysis Using Dominant Failure Modes Identified by Selective Searching
Technique.” Reliability Engineering & System Safety 119, pp. 316–31.
[21] Nakamura, S., and K. Suzumura. 2012. “Experimental Study on Fatigue
Strength of Corroded Bridge Wires.” Journal of Bridge Engineering 18,
no. 3, 200–09. 10.1061/(ASCE)BE.1943-5592.0000366
140 • BRIDGE RELIABILITY AND SERVICEABILITY
[22] Faber, M.H., S. Engelund, and R. Rackwitz. 2003. “Aspects of Parallel Wire
Cable Reliability.” Structural Safety 25, no. 2, pp. 201–25.
[23] Yang, W., P. Yang, X. Li, and W. Feng. 2012. “Influence of Tensile Stress
on Corrosion Behaviour of High-strength Galvanized Steel Bridge Wires in
Simulated Acid Rain.” Materials and Corrosion 63, no. 5, pp. 401–07.
[24] Mahmoud, K.M. 2007. “Fracture Strength for a High Strength Steel Bridge
Cable Wire with a Surface Crack.” Theoretical and Applied Fracture
Mechanics 48, no. 2, pp. 152–60.
[25] Maljaars, J., and T. Vrouwenvelder. 2014. “Fatigue Failure Analysis of Stay
Cables with Initial Defects: Ewijk Bridge Case Study.” Structural Safety 51,
pp. 47–56.
[26] Freire, A.M.S., J.H.O. Negrao, and A.V. Lopes. 2006. “Geometrical
Nonlinearities on the Static Analysis of Highly Flexible Steel Cable-stayed
Bridges.” Computers & Structures 84, no. 31, pp. 2128–40.
[27] Yoo, H., H.S. Na, E.S. Choi, and D.H. Choi. 2010. “Stability Evaluation of Steel
Girder Members in Long-span Cable-stayed Bridges by Member-based Stability
Concept.” International Journal of Steel Structures 10, no. 4, pp. 395–410.
[28] Yoo, H., H.S. Na, and D.H. Choi. 2012. “Approximate Method for Estimation
of Collapse Loads of Steel Cable-stayed Bridges.” Journal of Constructional
Steel Research 72, pp. 143–54.
[29] Liu, Y., N. Lu, M. Noori, and X. Yin. 2014. “System Reliability-based
Optimization for Truss Structures Using Genetic Algorithm and Neural
Network.” Int. J. Relia. Safe. 8, no. 1, pp. 51–69.
[30] Liu, Y., N. Lu, and X. Yin. 2016a. “A Hybrid Method for Structural System
Reliability-based Design Optimization and its Application to Trusses.” Qual.
Reliab. Eng. Int. 32, no. 2, pp. 595–608.
[31] Dai, H., H. Zhang, and W. Wang. 2012. “A Support Vector Density-based
Importance Sampling for Reliability Assessment.” Reliability Engineering
& System Safety 106, pp. 86–93.
[32] CSRA [Computer software]. 2013. Complex Structural Reliability Analysis
V 1.0, Changsha University of Science and Technology, Changsha, China.
[33] Tang, Q.Y., and C.X. Zhang. 2013. “Data Processing System (DPS) Software
with Experimental Design, Statistical Analysis and Data Mining Developed
for Use in Entomological Research.” Insect Science 20, no. 2, pp. 254–60.
[34] Chang, C.C., and C.J. Lin. 2011. “LIBSVM: A Library for Support Vector
Machines.” ACM Transactions on Intelligent Systems and Technology (TIST)
2, no. 3, p. 27.
[35] Ministry of Communications and Transportation (MOCAT). 2010. Stay
Cables of Parallel Steel Wires for Large-span Cable-stayed Bridge JT/T 775-
2010. Beijing, China,
[36] Ministry of Housing and Urban-Rural Development (MOHURD). 1999.
Unified Standard for Reliability Design of Highway Engineering Structures
GB/T 50283-1999. Beijing, China.
About the Authors
L computational framework,
LEFM. See Linear elastic fracture 72–74
mechanics Rice’s level-crossing formula,
Lifetime maximum deflection 70–72
assessment, 111–114 on two cable-supported bridges
Limit state function, 25–27, 36 probabilistic estimation,
Linear elastic fracture mechanics 77–81
(LEFM), 8 stochastic traffic load
Long-span bridges simulation, WIM
case study measurements, 74–76
extreme deflection, dynamic Probability density functions
traffic loads, 106–111 (PDFs), 18, 35
lifetime maximum deflection
assessment, 111–114 R
weigh-in-motion Rice formula, 93–95
measurements, 105–106 Rice’s extrapolation accounting,
deflection extrapolation, 96–98
102–104 Rice’s level-crossing formula,
dynamic performance of, 4–6 70–72
serviceability of, 89 Rice’s level crossing theory, 50–52
traffic loading behavior, 90–91 Road-roughness coefficient (RRC),
Long-span cable-stayed bridge, 93
132–137 RRC. See Road-roughness
coefficient
M
Monte Carlo simulation S
for long-span bridges, 4–5 SHM. See Structural health
stochastic traffic load simulation, monitoring
48–50 Short-span cable-stayed bridge,
128–132
N Simplified traffic-bridge
Nonstationary traffic load effects, interaction formulation, 69–70
90 Simulated stochastic fatigue truck
load model, 19–20
P S-N (stress-life) curve approach,
Parallel wire cable strength model, 25–26
121–122 Steel box-girder bridges, 7–8
Parametric studies, 82–84 Steel girders, 13
PDFs. See Probability density Stochastic fatigue truck load
functions model
Probabilistic analysis, 30–36 currently used, 16–17
Probabilistic modeling, 23–24 in design specifications, 16–17
first-passage probability, 54–57 proposed, 17–20
traffic load effects, 54–57 simulated, 19–20
146 • Index
Momentum Press is one of the leading book publishers in the field of engineering,
mathematics, health, and applied sciences. Momentum Press offers over 30 collections,
including Aerospace, Biomedical, Civil, Environmental, Nanomaterials, Geotechnical,
and many others.
Momentum Press is actively seeking collection editors as well as authors. For more
information about becoming an MP author or collection editor, please visit
http://www.momentumpress.net/contact
The Momentum Press digital library is very affordable, with no obligation to buy in future years.
For more information, please visit www.momentumpress.net/library or to set up a trial in the US,
please contact [email protected].
EBOOKS Multi-scale Reliability and
LU • NOORI
FOR THE SUSTAINABLE STRUCTURAL
ENGINEERING
Serviceability Assessment of SYSTEMS COLLECTION
LIBRARY In-service Long-span Bridges Mohammad Noori, Editor
Create your own
Naiwei Lu • Mohammad Noori
Customized Content
Bundle—the more With the development in global economic and transportation
books you buy, engineering, the traffic loads on brides have been g
rowing
Multi-scale
heavy traffic volumes and simultaneous truck loads on the
THE CONTENT bridge deck, and thus the safety and serviceability of the bridge
• Manufacturing deserves investigation. In this book, a multiscale reliability
Engineering
• Mechanical
method is presented for the safety assessment of long-span
bridges. The multiscale failure condition of stiffness girders Reliability and
Serviceability
& Chemical is the first-passage criteria for the large-scale model and the
Engineering fatigue damage criteria for the small-scale model.
• Materials Science
& Engineering
• Civil &
Environmental
It is the objective of this book to provide a more in-depth
understanding of the vehicle-bridge interaction from the
random vibration perspective. This book is suitable for
Assessment
Engineering
• Advanced Energy
adoption as a text book or a reference book in an advanced
structural reliability analysis course. Furthermore, this book also
of In-service
Long-span
Technologies provides a theoretical foundation for better understanding of
the safety assessment, operation management, maintenance
THE TERMS and reinforcement for long-span bridges and motivates further
• Perpetual access for
a one time fee
• No subscriptions or
research and development for more advanced reliability and
serviceability assessment techniques for long-span bridges. Bridges
access fees Naiwei Lu is a lecturer of civil engineering at Changsha
• Unlimited University of Science and Technology.
concurrent usage
• Downloadable PDFs Mohammad Noori is professor of mechanical engineering and
• Free MARC records ASME fellow in the department of mechanical engineering,
ISBN: 978-1-94708-338-7