Mangalathusivasubramanianpillai Dissertation 2017
Mangalathusivasubramanianpillai Dissertation 2017
A Dissertation
Presented to
The Academic Faculty
by
Sujith Mangalathu
In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy in the
School of Civil and Environmental Engineering
Approved by:
&
To those who are working hard to make the world a better place
ACKNOWLEDGEMENTS
This dissertation has been made with the help of many individuals and I am grateful for
their contribution and support. I thank my primary advisor Prof. Reginald DesRoches, who
granted me the privilege to work in his research group. I am grateful to him for the
discussions and advice that helped me sort out the technical details of my work. I also thank
my co-advisor Prof. Jamie Padgett for her comments and suggestions.
I would like to thank the support I have received from the California Department of
Transportation to pursue my doctoral studies. I would like to express my sincere gratitude
to Cliff Roblee, for his time and patience in providing insight regarding the bridge
inventory in California
I would like to thank my thesis committee Dr. Brani Vidakovic, Dr. Lauren Stewart and
Dr. Iris Tien. My special thanks to Dr. Vidakovic considering the many hours I spent in his
office and his lectures discussing probability theory and statistics. With his enthusiasm, his
inspiration, and his great efforts to explain things clearly and simply, he helped to make
statistics fun for me. I also thank Prof. Leonardo Osario (Rice University) for his
professional guidance. He provided encouragement, sound advice, good company, and lots
of good ideas.
My stay at Georgia Tech was a journey of learning and inspiration. The technical
discussion I had with fellow students Farahnaz Soleimani, Parsa Banihashemi, Edwin Lim,
Stephen Hsu, and Liu Xi helped me in the dissemination of knowledge in the subject and I
thank them for sharing their time. I will always remember the great conversations we had
over lunch, and I know that we will break bread together many more times. I also wish to
acknowledge Ellen Cormack for her motherly care and advice. Her door is always open for
me, and she always has time for my concerns.
I would like to thank the former research members of Prof. DesRoches group – Dr. Karthik
Ramanathan and Dr. Jong-Su Jeon. I admire the patience of Dr. Jeon in teaching me the
OpenSees and bridge design philosophies. I appreciate his help in writing research papers,
sharing his knowledge, and his active interest in my research. I also thank my friend Jiqing
Jiang for her support and discussions. I also thank her for helping me to not to lose
confidence and pointing the need for hard work.
I thank my friends Brian Terranova, Ni Pengpeng, Adris Ayala, Yasin Ogras, Jenni Tipler,
Yesudas, Gokulnath, Ramanandan, Ashkan, Anver Hisham, Jayakrishnan, Andrea,
Dharanidharan, Kiran, Anoob, Yesudas, Prabhash, Nikhil, Gayathri for all the
encouragement and support. I believe that I am tremendously fortunate to establish the
friendship with these outstanding people. I appreciate Brian, Ni, and Gokulnath for helping
me get through the difficult times, and for all the emotional support, entertainment, and
caring they provided. I am not sure I would be at Georgia Tech without their
encouragement.
I also thank my friends and mentors Dr. Vipin Unnithan, Dr. Robin Davis and Shinto Paul
for their inspiration and guidance. The conversations, advice, friendships, and fun helped
shape my development as a researcher and a good human being.
I thank my non-Georgia tech friends at Atlanta, Karthika, Arun, Vinod, Indu, Gireesh,
Kutty, Mahesh and Subin for the countless fun and activities. They helped me in numerous
ways which are impossible to describe. It’s not easy to forget the hiking trips and poker
games we had over weekends.
I would be remiss if I didn’t acknowledge the contribution and support I have received
from Surya. Over the years she has expressed confidence in my abilities and supported me
unconditionally. My love and appreciation is extended to her.
I am deeply indebted towards the movies for providing me a recreational and artistic mind
during my Ph.D, particularly the movies of Christopher Nolan, Martin Scorsese, Giuseppe
Tornatore, Majid Majidi, Ki-Duk Kim, Asghar Farhadi, Ang Lee and Clint Eastwood. The
movies - The Lives of Others, Life is Beautiful, Cinema Paradiso, Hotel Rwanda, The
Prestige - has always a special place in my heart.
I also acknowledge the legendary soccer players Pele, Maradona and Messi for entertaining
and motivating me since childhood. I also acknowledge my soccer colleagues at Georgia
Tech.
I also extend my thanks to the dedicated and silent service of many at Georgia Tech
(Georgia Tech Police, CRC Staff, administrative staff, etc.). Their service helps my stay at
Georgia Tech a nice and pleasant one.
I also appreciate the dedicated and patient service of people around the globe to make the
world a better place.
As the adage goes - like any journey, it's not what you carry but what you leave behind.
Disclaimer
The current study has been supported by the California Department of Transportation
(Caltrans), USA through Project P266, Task 1780: Production development of generation-
2 fragility models for California bridges. Any opinions, findings, and conclusions or
recommendations expressed in this thesis are those of the author and do not necessarily
ACKNOWLEDGEMENTS iv
LIST OF TABLES vii
LIST OF FIGURES ix
LIST OF SYMBOLS AND ABBREVIATIONS xi
SUMMARY xv
CHAPTER 1 INTRODUCTION 1
1.1 Problem Description and Motivation 1
1.2 Research Objectives 5
1.3 Dissertation outline 6
CHAPTER 2 EXISTING RESEARCH ON GROUPING OF BRIDGE CLASSES
AND BRIDGE FRAGILITY 8
2.1 Grouping of bridge classes 9
2.2 Fragility curves 12
2.2.1 Expert opinion 12
2.2.2 Empirical methods 12
2.2.3 Analytical methods 13
2.2.4 Parameterized fragility curves 16
2.3 Fragility curves for concrete bridge classes in California 19
2.4 The need to go beyond HAZUS-based grouping and fragility curves 22
2.5 Uncertainty treatment in fragility analysis 27
2.6 Closure 32
CHAPTER 3 MODELING OF BRIDGE COMPONENTS 34
3.1 Superstructure 35
3.2 Substructure 36
3.2.1 Bents 36
3.2.2 Columns 37
3.2.3 Idealization of bridge columns 43
3.2.4 Validation of bridge columns 47
3.3 Abutments 47
3.4 Bearings 52
3.5 Shear Keys 53
3.6 Pounding 54
3.7 Foundation 55
3.8 Closure 57
CHAPTER 4 PERFORMANCE BASED GROUPING OF BRIDGE CLASSES 59
4.1 Review of ANOVA, ANCOVA, and Kruskal–Wallis approach 60
4.2 ANOVA based grouping 61
4.3 ANCOVA based grouping 63
4.4 KW-based grouping 65
4.5 Case Study: Two-and Three-Span Box-Girder Bridges in California 66
iv
4.6 Comparison of various grouping techniques 72
4.6.1 Case 1: Significant per ANCOVA 72
4.6.2 Case 2: Significant per ANOVA and ANCOVA 75
4.6.3 Case 3: Significant per KW 76
4.7 Grouping of bridge classes 79
4.8 Conclusion 86
CHAPTER 5 CALIFORNIA BRIDGE INVENTORY 88
5.1 Bridge classification based on BIRIS 89
5.2 Box girder bridge class statistics 89
5.3 Abutments 90
5.4 Bearings 95
5.5 Box girder deck 96
5.5.1 Span length 96
5.5.2 Deck width 97
5.5.3 Deck cross-section properties 100
5.6 Columns 102
5.6.1 Column height 102
5.6.2 Column cross-section 103
5.6.3 Column material properties 106
5.6.4 Column reinforcement details 106
5.7 Foundations 107
5.8 Other uncertain parameters 111
5.8.1 Damping 111
5.8.2 Mass factor 111
5.8.3 Shear key acceleration 111
5.8.4 Gap 112
5.8.5 Earthquake direction 112
5.9 Closure 112
CHAPTER 6 SYSTEM AND COMPONENT FRAGILITY CURVES FOR BOX-
GIRDER BRIDGES 114
6.1 Fragility Framework 114
6.1.1 Probabilistic seismic demand models 116
6.1.2 Capacity models 117
6.1.3 Abutments 128
6.2 Fragility methodology 130
6.2.1 Ground motion suite 133
6.2.2 Material and geometric uncertainties and parameterized stochastic bridge
models 134
6.3 Fragility curves for multi-span continuous concrete single frame box-girder
bridges 134
6.3.1 Trends based on design era 140
6.3.2 Trends based on spans 142
6.3.3 Trends based on abutment type 143
6.3.4 Trends based on column cross-section 144
6.3.5 Trends based on number of columns per bent 146
6.4 HAZUS comparison 147
v
6.5 Closure 152
CHAPTER 7 PARAMETERIZED FRAGILITY CURVES: LASSO APPROACH
154
7.1 Regression models 156
7.1.1 Linear Regression 156
7.1.2 Stepwise regression 157
7.1.3 Ridge regression 157
7.1.4 Lasso regression 158
7.1.5 Elastic net 159
7.2 Case-study bridges: numerical modeling, uncertainties, ground motion suite,
and demand parameters 159
7.3 Comparison of the regression models 163
7.3.1 Investigation of penalty factor 164
7.3.2 Comparison of the regression models 167
7.4 Sensitivity of input parameters to the seismic demand model 170
7.5 Multi-Parameter fragility curves 172
7.6 Conclusions 179
CHAPTER 8 CONCLUSIONS AND FUTURE WORK 181
8.1 Summary and Conclusions 181
8.2 Research Impact 185
8.3 Recommendations for future work 187
APPENDIX A. EARTHQUAKE RECORDS USED FOR FRAGILITY ANALYSIS
189
APPENDIX B. FRAGILITY CURVES IN TERMS OF PGA 207
APPENDIX C. COMPONENT FRAGILITY CURVES FOR BRIDGE CLASSES
210
REFERENCES 230
vi
LIST OF TABLES
Table 2.1 HAZUS grouping and fragility relationships for bridge classes in California. 23
Table 3.1 – Idealization of bridge columns. 44
Table 4.1 – Uncertainty distribution considered in the bridge models 70
Table 4.2 – p–values by ANCOVA, ANOVA and KW test 71
Table 4.3 – p–values from ANOVA. 80
Table 4.4 – Results of the grouping for two-span box girder bridges. 82
Table 4.5 – Results of the grouping for multi-span bridges 83
Table 5.1 – Bridge classes in California and their proportion in the overall inventory. 89
Table 5.2 – Distribution of parameters for abutments. 94
Table 5.3 – Distribution of parameters for bearings. 95
Table 5.4 – Distribution of span length and span ratio (approach span/main span) for box
girder bridges. 98
Table 5.5 – Distribution of deck width for box girder bridges. 99
Table 5.6 – Deck cross-section properties. 100
Table 5.7 – Box-girder deck slab thickness (MTD, 2008). 101
Table 5.8 – Column height distribution. 103
Table 5.9 – Distribution of column cross-sections. 104
Table 5.10 – Statistical distribution of column material properties. 106
Table 5.11 – Statistical distribution of column reinforcement details. 107
Table 5.12 – Distribution of foundation rotational stiffness (10 kip-in/rad).
6
109
Table 5.13 – Distribution of foundation translational stiffness (kip/in). 110
Table 5.14 – Distribution of other uncertain parameters. 112
Table 6.1 – Engineering demand parameters for bridge components monitored in NLTHA.
116
Table 6.2– Component level damage state descriptions – Component Damage Thresholds
(CDT). 118
Table 6.3 – General description of BSSTs along with CDTs. 119
Table 6.4 – General definition of column capacity limit states. 120
Table 6.5 – Design details for columns in Era 11 (pre-1971). 122
Table 6.6 – Design details of columns in Era 22 (1971-1990). 123
Table 6.7 – Design details for columns in Era 33 (post-1990). 124
Table 6.8 – Summary of limit states for Era 11 columns. 125
Table 6.9 – Summary of limit states for Era 22 columns. 126
Table 6.10 – Summary of limit states for Era 33 columns. 127
Table 6.11– Statistical summary of ductility values for various design eras. 128
Table 6.12– Median value of CDT for abutment seat. 129
Table 6.13– Summary of CDT values for various bridge components. 130
Table 6.14 – Nomenclature adopted in the current study. 135
Table 6.15 – Fragility values for two span continuous concrete box-girder fragilities with
diaphragm abutments. 137
Table 6.16 – Fragility values for two span continuous concrete box-girder fragilities with
seat abutments. 137
vii
Table 6.17 – Fragility values for multi-span continuous concrete box-girder fragilities with
diaphragm abutments. 138
Table 6.18 – Fragility values for multi-span continuous concrete box-girder fragilities with
seat abutments. 138
Table 6.19 – Comparison of bridge classes. 148
Table 7.1 – Uncertainty distribution considered in the bridge models. 162
Table 7.2 – Bridge component demand parameters. 163
Table 7.3 – Estimated coefficients and test error for COL for the bridge with diaphragm
abutments by various regression techniques 168
Table 7.4 – Limit state models of various bridge components. 175
viii
LIST OF FIGURES
ix
Figure 4.8 – Proposed classification scheme. 83
Figure 5.1 – Illustration of major bridge components. 90
Figure 5.2 – Distribution of abutments for various eras. 91
Figure 5.3 – Distribution of abutments for various bridge eras. 92
Figure 5.4 – Percentage distribution of bearings based on design eras. 95
Figure 5.5 – Cross-section details of box-girder bridges. 100
Figure 5.6 – Column cross-sections for various design eras of box girder bridges. 105
Figure 5.7 – Bridge foundation types (Priestley et al. 1996). 107
Figure 5.8 – Statistical distribution of foundation for various eras. 108
Figure 6.1 – Schematic representation of the NLTHA procedure used to derive the
PSDMs. 115
Figure 6.2 – Illustration of a typical PDSM. 117
Figure 6.3 – Schematic of the fragility framework. 132
Figure 6.4 – Response spectra for the selected ground motions. 133
Figure 6.5 – Illustration of change in median value and relative vulnerability. 136
Figure 6.6 – System and component fragility curves for bridge classes S-E1-S22-C-D
and S-E2-S22-C-S. 140
Figure 6.7 – Plot of median values of bridge classes across the design eras for a) S-
E(1/2/3)-S22-C-D, and b) T-E(1/2/3)-S34-C-S. 141
Figure 6.8 – Plot of median values of bridge classes across the number of spans for
bridge classes S-E2-S(22/34)-C-D and S-E2-S(22/34)-O-D. 143
Figure 6.9 – Plot of median values of bridge classes across the type of abutments for
the selected bridge classes. 143
Figure 6.10 – Plot of median values of bridge classes across various column cross-
sections for the selected bridge classes. 145
Figure 6.11 – Plot of median values of bridge classes across the number of columns
per bent for the selected bridge classes with a) diaphragm abutments and b) seat
abutments. 147
Figure 6.12– Comparison of HAZUS and selected bridge class fragilities for a) Era 11
bridges with single column bent and b) Era 22 and Era 33 bridges with single column
bent. 150
Figure 6.13– Comparison of HAZUS and selected bridge class fragilities a) Era 11
two-span bridges with two- and multi-column bents and b) Era 22 and Era 33 two-
span bridges with single and multi-column bents. 151
Figure 7.1 – Numerical modeling of various bridge components. 161
Figure 7.2 – Comparison of MSE with . 166
Figure 7.3 – Shrinkage of regression coefficients with . 167
Figure 7.4 – Radar plot depicting the comparison of accuracy of fit obtained from the
various regression models. 170
Figure 7.5 – Sensitivity of input parameters. 172
Figure 7.6 – System and component fragility curves for moderate damage state: a)
diaphragm abutment bridge, b) seat abutment bridge. 176
Figure 7.7 – Sensitivity of fragility curves to input parameters for diaphragm abutment
bridge for various limit states 177
Figure 7.8 – Sensitivity of fragility curves to input parameters for seat abutment
bridge for various limit states 178
x
LIST OF SYMBOLS AND ABBREVIATIONS
xi
fcc maximum concrete stress
fyh yield strength of the transverse reinforcement
G(EDP|IM) CDF of EDP conditioned on IM
G(DM|EDP) CDF of DM conditioned on EDP
G(DV|DM) CDF of DV conditioned on DM
g acceleration due to gravity
H column height
H0 null hypothesis
H1 alternate hypothesis
Ha abutment height
Hal alternate hypothesis
JPSDM joint probabilistic seismic demand model
IDA incremental dynamic analysis
IM ground motion intensity
KW Kruskal-Wallis test
Kpa stiffness of abutment piles
Kb stiffness of bearing
Kfr rotational stiffness of foundation piles
Kft translational stiffness of foundation piles
ke effective confinement factor
L span length
Lc width of rectangular/oblong cross-section
LN log-normal distribution
M multi column bent
MSE mean square estimate
m mass factor
N total sample size/number of simulations
NLTHA nonlinear time history analysis
NSP nonlinear static procedures
O oblong cross-section
PC pre-stressed concrete
PGA peak ground acceleration
PSDM probabilistic seismic demand model
R rectangular cross-section
RC reinforced concrete
S single column bent
S11 single span bridge
S22 two span bridge
S34 three and four span bridge
S5x bridges with more than five spans
Sa-1.0s spectral acceleration at a period of 1.0 s
Sd median estimate of the demand as a function of the IM
Sc median estimate of the capacity
s center to center spacing or pitch of spiral or circular hoop
SDOF single-degree of freedom
SST total sum of squares
xii
T two column bent
UST superstructure unseating displacement
(•) standard normal cumulative distribution function
displacement
t gap between the deck and shear key
regression coefficients
d|IM dispersion of the demand conditioned on the IM
a abutment displacement in active direction
b bearing displacement
d deck displacement
fnd foundation displacement
p abutment displacement in passive direction
seat seat displacement
t abutment displacement in transverse direction
u unseating displacement
key shear key displacement
y yield displacement
cc maximum concrete strain
penalty factor
mean of fragility curves
μ mean value
μas coefficient of friction for abutments on spread footing
μb coefficient of friction of bearing pad
μ column curvature ductility
DV mean annual frequency of a decision variable
longitudinal reinforcement ratio
cc ratio of area of longitudinal reinforcement to the area of core
of section
f foundation rotation
σ standard deviation
damping ratio
dispersion of fragility curves
2SSC two span seat-abutment bridge with circular cross section and
single-column bent
2SSR two span seat-abutment bridge with rectangular cross section
and single-column bent
2SSS two span seat-abutment bridge with abutments resting on
spread footing
2SSS two span seat-abutment bridge with abutments resting on
piles
2SDPS two span diaphragm-abutment single-column bent bridge
with pre-stressed concrete superstructure
2SDRS two span diaphragm-abutment single-column bent bridge
with reinforced concrete superstructure
xiii
3SDC three span seat-abutment bridge with circular cross section
and single-column bent
3SDR three span seat-abutment bridge with rectangular cross
section and single-column bent
3SCMCB three span seat-abutment bridge with rectangular cross
section and multi-column bent
3SRMCB three span seat-abutment bridge with circular cross section
multi-column bent bridge
xiv
SUMMARY
Fragility curves play a critical role in regional seismic risk assessment and are a key
to provide fragility curves that best represent the bridge inventory. However, it is
impractical to develop unique fragility curves for each structure across a regional portfolio.
One strategy that has been adopted to address this challenge is to group bridges into classes
with similar design or structural performance. Traditionally, this grouping has been
identification leads to a number of bridge classes and unwarranted grouping. This work
California, and is the first systematic approach in sub-binning bridge classes for the
regional risk assessment. The proposed grouping and analytical fragility methodology is
used to derive fragility relationships for single frame box girder bridges in California. This
work concludes with the application of machine learning techniques for the generation of
xv
CHAPTER 1 INTRODUCTION
recent earthquakes have demonstrated that these bridges are one of the most vulnerable
serious threat to immediate recovery efforts and cause communities to incur large
economic losses. Recovery efforts and loss estimation can typically be calculated via
damage functions of structures called fragility functions (Kircher et al., 2006). A fragility
curve is defined as a conditional probability that gives the likelihood that a structure or
component of that structure will meet or exceed a certain level of damage for a given
ground motion intensity (IM). This enables the realistic estimation of economic losses, as
well as planning for emergency responses and ascertaining the need for retrofitting.
approaches (Basöz and Mander, 1999; Banerjee and Shinozuka, 2007; Gardoni et al., 2003;
Zhong et al., 2008; Nielson 2005; Mangalathu et al. 2015a). Empirical methods have been
used to develop fragility curves in regions where extensive earthquake records are
available. Empirical curves are based on observed damage from previous earthquakes.
These types of fragility curves tend to be the most realistic, but are very specific to a
particular earthquake and structure and thus have limited application (Jeong and Elnashai,
2007). The limitations of the empirical approach motivate the generation of fragility curves
using analytical approaches. Analytical fragility curves are generated using numerical
simulations that account for material, geometric, and ground motion uncertainties.
1
California is a state with a high seismic hazard and a history of damaging
Stojadinović, 2006; Zhang and Huo, 2009) have developed fragility curves for bridges in
California that have proved to be valuable in understanding the behavior and seismic
vulnerability of bridges. However, the developed fragility curves are structure-specific, and
therefore are only beneficial for the risk assessment of a specific bridge (Ramanathan et
al., 2015). Fragility curves proposed by Basöz and Mander (1999) are remotely applicable
to bridge classes in California and are used in HAZUS (HAZUS-MH, 2003). However, the
analysis, a limited number of bridge parameters, and damage states based on a limited set
of field observations. The effects of interior supports, framing systems, and design
This issue was partially addressed in the work of Ramanathan et al. (2015) by
generating fragility curves that are applicable to a portfolio of bridges in California. These
curves considered uncertainties in attributes such as span length, column height, number
of spans, superstructure type, and material properties. However, a detailed review of bridge
plans from in-house databases obtained from the California Department of Transportation
(Caltrans) shows that their study addresses only a specific class of bridges with a specific
column shape (circular), bearing type (elastomeric), and abutment type (abutment on piles).
To approximately cover the entire range of the California bridge inventory, their study
extension supports regional risk assessment for transportation networks and helps agencies
2
The proposed research strives for the generation of fragility curves for various
classes of box-girder concrete bridges in California. Each class of bridge systems relies on
section, abutment type, type of interior support (number of columns per bent), and design
era. Such a classification leads to numerous subclasses for a particular bridge type, which
makes it cumbersome to compute the seismic fragility for each subclass. It is also not clear
whether all such combinations can yield distinct bridge performance classes, or which
Additionally, it is not yet clear whether such a detailed classification would result in a better
refinement of the vulnerability assessment. The initial study is directed towards the
Various studies (Saiidi et al., 1996; Jangid, 2004; Nielson and DesRoches, 2006;
Padgett and DesRoches, 2007) have been conducted to evaluate the sensitivity of fragility
curves to various input parameters. However, there is a lack of understanding of the effects
of various bridge attributes on the fragility curves and the necessity of this understanding
to creating groups of bridge classes. Mangalathu et al. (2015b) addressed this issue through
Probabilistic Seismic Demand Model (PSDM). The research carried out underscored the
attributes on fragility curves. However, their study is limited to two- and three-span single
frame bridge configurations constructed before 1970. The effects of span, design era,
number of frames, and other attributes have not been addressed in their study. The grouping
3
of their work is addressed in this study through a simple method based on Analysis of
means, and a grouping methodology based on ANOVA is suggested in this current study.
The methodology also helps to identify which bridge attributes significantly impact the
seismic bridge response and hence fragility curves. The proposed approach and the
inventory Representative bridge systems (called RBS- hereafter) so that it can fairly
represent fairly the entire bridge inventory. Real bridges in the California inventory are
conform well to one of the proposed RBS classes, many real bridges have unique design
features and/or combinations that do not ‘fit’ neatly within the proposed RBS class
definitions. To improve fit (or modeling fidelity), one could add a new RBS class to directly
capture the unique features but only at the cost of increasing the number of RBS classes
requiring analysis. At the extreme, this strategy would lead to the development of bridge-
specific models, which is not warranted and is beyond the scope of this work. The current
study generates fragility curves for the single frame box-girder bridge classes using the
fragility relationships for single frame box-girder bridges in California. The study is limited
to straight bridges and the effects of skew, curvature, and unbalanced frame will not be
addressed.
4
1.2 Research Objectives
The limitation in the HAZUS grouping of bridge classes and the corresponding
fragilities are identified in the subsequent sections. This research aims to improve the
• Identify whether it is rational to go beyond the existing HAZUS grouping and fragility
relationships.
grouping) to group bridge classes with statistically similar performance and damage
measures. The proposed methodology will account for the effects of design eras, cross-
sections, number of spans, number of frames, abutment types, span continuity, and pier
• Perform a detailed plan review of various bridge classes in Californian through review
engineers.
• Generate statistically significant yet nominally identical bridge models accounting for
5
• Generate a refined set of fragility curves at system and component level for various
box-girder bridge classes in California. These improved fragility curves will help
determine the relative vulnerability of various bridge classes. They will also assist the
using advanced statistical and machine learning techniques such as Lasso, Ridge, and
elastic net. Such study will provide insight in quantifying whether the variation of
those parameters which have minimal influence on the seismic demand and reduces
The research is organized into seven subsequent chapters with the following contents:
Chapter 2 presents an overview of existing literature on the grouping of bridge classes and
Chapter 3 provides extensive details about the modeling strategies of various bridge
The chapter also presents the integration of various component models to generate global
6
Wallis (KW) towards the grouping of structures of similar performance. The chapter also
Chapter 5 presents an in-depth study of the California bridge inventory using the in-house
database called BIRIS, assembled by Caltrans engineers. The chapter also presents
fragility curves for box girder bridges in California. The chapter also describes the system
and component fragility curves for single frame multi-span box girder bridges in
California. Insights are provided on the relative performance of various bridge classes and
Chapter 7 explores the application of regression and machine learning techniques for the
Chapter 8 presents the conclusions from the present research, key contributions and the
7
CHAPTER 2 EXISTING RESEARCH ON GROUPING OF
framework (Cornell and Krawinkler, 2000; Moehle and Deierlein, 2004) has evolved to be
the next-generation framework in risk mitigation decision making for structure and
Research Center (PEER) is the widely accepted robust methodology for PBEE. The PEER
rigorous treatment of uncertainties. The underlying approach is shown in Figure 2.1, and
the framework assumes that the performance assessment of components as discrete Markov
process, where the conditional probabilities between parameters are independent (Moehle
8
where DV is the mean annual frequency of a decision variable (DV, e.g., repair cost,
demand model describing the CDF of EDP conditioned on ground motion intensity
measure (IM, e.g., peak ground acceleration), and (IM) is the seismic hazard model
G(DM|EDP) and G(EDP|IM) yields fragility curves. A fragility curve can thus be defined
as a conditional probability that gives the likelihood that a structure or component will
meet or exceed a certain level of damage for a given ground motion intensity (IM). The
fragility curves intended for regional risk assessment require grouping of bridge classes
Any existing bridge has its own structural characteristics due to its location, soil
bridges with similar structural properties are expected to show statistically similar
performance under a given earthquake loading, and so the bridges with similar performance
can be grouped together. The existing literature on the grouping of bridge classes is given
in this section.
grouping the bridge classes and seismic vulnerability estimation. HAZUS grouped the
bridge classes with similar damage/loss characteristics and suggested fragility relationships
to the grouped bridge classes. HAZUS classified the bridge classes based on seismic
design, number of spans, span length, bent type, span continuity, and span discontinuity,
9
and is shown in Figure 2.2. The HAZUS grouping was based on engineering judgment,
past experience, and expert opinion. The effects of evolution in seismic design philosophy,
column cross-section, and number of frames are not addressed in HAZUS. The limitations
column circular, single column rectangular, multi-column, or wall-type), deck type (slab,
combinations are possible, the authors further reduced the bridge classes to 11 classes
(those with five or more bridges). Based on past earthquake data and the previous research,
Avsar et al. (2011) classified the highway bridges in Turkey that were constructed after
1990. The important structural attributes identified by the authors are span number (single
or multiple), bent (single or multiple), and skew angle (negligible or significant, chosen to
be >30).
Ramanathan et al. (2015) grouped the bridge classes in California based on limited
columns per bent (single or multiple), superstructure type (box girder, I-girder, T-girder,
10
or slab bridges) and design era (pre-1971, 1971-1990, or post-1990). However, the
grouping was based on engineering judgment and doesn’t consider the effects of number
of spans, span discontinuity, span length, or cross-section. Thus, the grouping covers only
the probabilistic seismic demand models (PSDMs) of different bridge classes. A PSDM is
defined as the probability distribution of structural demands (D) conditioned on the ground
motion intensity measure (IM). This works presents the first systematic and reliable
methodology for grouping bridge classes for performing regional risk assessments. The
authors demonstrated their grouping methodology through case studies of two-span and
three-span box-girder bridges in California with various design attributes. Their research
showed the importance of binning of bridge classes through the comparison of fragility
curves for bridge classes with different design attributes as identified by the ANCOVA.
However, the focus of the grouping was only on two-span and three-span box-girder
bridges and doesn’t consider the effect of number of spans, foundation type, design era, or
span discontinuity.
Mehr and Zaghi (2016) used analysis of variance (ANOVA) to compare the
response of single-frame and multi-frame bridges and group them accordingly. The authors
investigated the effects of number of frames, soil type, substructure system, valley shape,
intensity of ground motion, and design capacity-to-demand ratio on single and multi-frame
bridges. Using the ANOVA results, the authors concluded that a multi-frame system is
more robust than a single frame system from a seismic perspective. The authors also noted
11
The existing bridge groupings suffer some limitations and cannot be used to group
the California bridge inventory. Also the structural attributes chosen to classify the bridges
vary depending on the type of bridge, bridge location, and research intention. California
has close to 29,000 bridges, which vary in age. In order to obtain a reliable estimate of the
risk associated with the California bridge inventory, it is crucial to group the entire bridge
class that yields similar performance or suffers similar damage following a seismic event.
Past decades have witnessed the development of several fragility curve generation
The earliest attempt to develop a fragility curve was based on expert opinion (ATC,
estimates and the results were presented in the form of damage probability matrices, later
converted to vulnerability functions and restoration curves. However, the technique was
wholly subjective and depended on the number of experts queried. The limitations of
expert-opinion fragility curves, coupled with actual damage data from earthquakes,
motivated the generation of empirical fragility curves (Basoz and Kiremidjian, 1998;
Basoz and Kiremidjian (1998) assembled data regarding damage to bridges from
the 1989 Loma Preita and the 1995 Northridge earthquakes in California and analyzed
them to obtain the relationships between bridge damage and ground motion. The authors
12
Shinozuka et al. (2000) used bridge damage data from the 1995 Kobe earthquake and used
distribution describing the fragility curves. These types of fragility curves tend to be the
most realistic, but are very specific to a particular earthquake and structure and thus have
limited application (Jeong and Elnashai, 2007). The limitations of the empirical approach
Analytical methods can be used to generate fragility curves where earthquake data
is not available. Various researchers have employed analysis techniques, with varying
and Hwang et al. (2000) extended this approach by quantifying the uncertainties in capacity
and demand assessments. With this advancement in the modeling capabilities, researchers
moved to nonlinear static procedures (NSP). Capacity spectrum method (CSM) and N2
method are the different types of NSP. Developed by Fajfar (2000), N2 method combines
the pushover analysis of a multi-degree of freedom (MDOF) system with the response
spectrum analysis of an equivalent SDOF system. HAZUS uses the fragility relationships
suggested by Mander and Basöz (1999), which are based on CSM. Further details on the
fundamental assumptions and limitation of HAZUS fragilities are given in the next section.
nonlinear time history analysis (NLTHA) (Banerjee and Shinozuka, 2007; Choi, 2002;
Gardoni et al., 2003; Kim and Shinozuka, 2004; Mackie and Stojadinovic, 2001; Mackie
and Stojadinović, 2005; Mangalathu et al., 2016a; Mangalathu et al., 2015; Padgett, 2007;
Ramanathan et al., 2015). Recalling the fragility curve as the probability that the seismic
demand (D) placed on a component exceeds the capacity (C), the probability can be
13
computed using Equation 2.1, assuming lognormal distribution for the D and C ( Cornell
et al., 2002)
ln( S / S )
P[ D C | IM] d c
(2.1)
d / IM c2
2
where Sd is the median estimate of the demand as a function of the IM, Sc is the
median estimate of the capacity, d|IM is the dispersion of the demand conditioned on the
IM, c is the dispersion of the capacity, and (•) is the standard normal cumulative
distribution function. Sd and d|IM can be estimated from the PSDMs. As previously
conditioned on the ground motion intensity measure (IM). NLTHA employs analysis of
bridges with different ground motion intensities to obtain the PSDMs. PSDMs can be
(Figure 2.4). The cloud approach consists of selecting ground motions that represent the
hazard at a region and carrying out NLTHA on the bridge samples. This technique is
In the IDA approach, ground motions are scaled successively until significant
reduction (collapse) of the primary load-bearing elements in the structural system. Hence,
IDA can offer the transition of the structural response from elastic to inelastic behavior,
finally leading to global dynamic instability, and the accurate and reliable estimates of the
global collapse capacity of the structure. The overall formulation of IDA was proposed by
(Vamvatsikos and Cornell, 2002). A significant drawback with this approach is that the
process involves scaling of ground motions without altering the frequency content of the
ground motions. These scaling approaches could lead to unrealistic time histories that
14
Figure 2.3 – Generation of PSDMs through cloud approach (Ramanathan, 2012).
15
Figure 2.4 – Generation of PSDMs through IDA approach (Ramanathan, 2012).
curves have some limitations: (1) the inability to account for the influence of uncertainty
16
(modeling) parameters on structural performance during earthquakes without extensive re-
simulations for each new set of parameter combinations; (2) the inability to explicitly
address the effect of uncertainty parameters on fragility curves; and (3) the inability to
fragility curves that can account for the variation in the design details or geometric
parameters of bridges (Seo and Linzell, 2012; Dukes, 2013; Ghosh et al., 2013; Kameshwar
and Padgett, 2014; Park and Towashiraporn 2014; Jeon et al., 2015; Mangalathu et al.,
2015; Mangalathu et al. 2017c). Assuming that the input variables are statistically
parameter) is constructed. Samples obtained from this demand model are compared with
those of the associated limit state model to obtain the binary survival-failure vector. This
coefficients and thus develop the multi-parameter fragility curve in the component.
Seo and Linzell (2013) used response surface models to generate parameterized
fragility curves for curved steel bridges. The authors identified the critical range of the
two example bridges using the response surface method and then derived parameterized
fragility curves for the bridges using logistic regression. The parameterized fragility curves
were used to produce bridge-specific fragilities by substituting a specific value for each of
six design parameters: longitudinal reinforcement ratio; volumetric transverse ratio; aspect
ratio; span length-to-column height ratio; deck depth-to-column diameter ratio; and deck
width. This framework was developed for use in the seismic design process in the design
17
of new bridges. It produces fragility curves without the need to create the curves
Ghosh et al. (2013) used a multi-parameter demand model to account for the effect
of all uncertain parameters in the generation of fragility curves. The authors used four
regression splines, radial basis function networks, and support vector machines) to
determine the best-fitting parameterized demand models involving the uncertain input
parameters. To achieve this goal, these authors selected for their case study multi-span
simply supported concrete bridges that were not seismically designed, which are typical in
the central and southeastern United States (CSUS). They then used ten parameters
associated with material and geometric uncertainties, along with an IM (eleven predictor
variables), to develop demand models. This work concluded that the MARS model
provided the most accurate estimates of component responses with the fewest predictive
errors. Using the MARS model and logistic regression, parameterized fragility curves were
hazard risk assessment for highway bridges subjected to earthquake and hurricane events.
The authors used stepwise logistic regression with a non-linear logit function to generate
the parameterized fragility curves. The significant parameters were identified by the
authors using a sequential forward selection scheme. The authors demonstrated the
proposed approach with the case studies on multi-span simply supported concrete bridges
in South Carolina.
number of spans, pier height, and earthquake magnitude. The authors used response surface
18
modeling to create second-degree polynomials for the estimation of seismic damage. The
study revealed that span length does not significantly affect seismic damage to bridges.
fragility estimates. The framework includes the selection of a bridge class, characterization
logistic regression and one-dimensional fragility curves using a Monte Carlo integration.
Additionally, the Bayesian approach used in the suggested framework enables the
performing numerous structural analyses required for design of experiments. The authors
demonstrated their approach through a study of two classes of curved bridges commonly
bridge-specific fragility curves in which the limit state of the bridges is explicitly defined,
accounting for the effects of varying geometry, material properties, reinforcement, and
loading patterns. The methodology can account for the uncertainty in capacity, demand,
and damage state definition. The authors used nonlinear static analysis and IDA to estimate
the demand and capacity, and reduced sampling techniques for the uncertainty treatment.
earthquakes. Various researchers have generated fragility curves for bridges in California,
using either an empirical (Başöz and Kiremidjian, 1996; Shinozuka et al., 2000) or
19
analytical approach. This section details the existing analytically based research on the
on seismic design, span length, bent type, and span discontinuity. However, HAZUS
fragility relationships were developed on the basis of a limited number of parameters and
simplified two-dimensional analysis, and did not account for the uncertainties in geometric
attributes for bridge classes such as the number of spans, span length, deck width, and
reflecting the variation in bridge design parameters, including the skewness, span length,
span to column height ratio, and column to superstructure dimension ratio. However, their
Ramanathan (2012) generated fragility curves that are applicable to a portfolio for
various classes of bridges by accounting for uncertainties in attributes such as span length,
column height, number of spans, superstructure type, and material properties. They
addressed the evolution in seismic design philosophy by grouping the bridge classes into
three eras: pre-1971 bridges (Era 11, hereafter), 1971-1990 bridges (Era 22, hereafter), and
post-1990 bridges (Era 33, hereafter). Although the study provides valuable insight
regarding the bridge fragilities, it has some limitations, which are noted below:
• The grouping of bridge classes in the study was carried out based on a traditional
subjective approach that relies on engineering judgment. Such subjective grouping has
20
• The study considers that only one type of abutment footing (abutment on piles) is
possible in California. However, the plan review of California bridges revealed various
abutment footings such as abutment on spread footing, abutment on piles, and tall
cantilever footing. Recent studies (Mangalathu et al., 2015; Mangalathu et al., 2016a)
have noted that bridge fragilities are significantly influenced by the abutment footing
type.
• The study was limited to specific bridge classes in the California bridge inventory. For
example, the study on box-girder bridges was limited to two-span bridges, yet bridges
• The study only addressed a specific class of bridges with circular column shape. It has
been noted from the plan review that various cross-sections such as circular,
rectangular, and oblong (interlocking spirals) are present in the California bridge
inventory.
• The study assumed that the bearings in seat abutments are elastomeric. However,
• The study considers only flexural mode of failure for columns in bridges constructed
during Era 11, although the lap-splice mode of failure is common in bridges from that
era.
bridges (Era 22 and Era 33) to flare the columns in the upper region to provide support
to the cap beam under eccentric live load for architectural reasons. The response of the
bridge columns to seismic loading is significantly affected by the flares (Sanchez et al.,
1997).
21
• The study didn’t consider the effects of frames, pier-type columns, and spread-type
• The capacity estimates or limit state models in Ramanathan (2012) were preliminary
HAZUS is the most comprehensive document for grouping bridge classes and
estimating seismic vulnerability. HAZUS groups bridge classes with similar damage/loss
characteristics and suggested fragility relationships to the grouped bridge classes. This
section summarizes the HAZUS grouping and fragility relations, and discusses their merits
and faults. Figure 2.2 shows the HAZUS-based grouping for the selected California box-
girder bridge inventory; Table 2.1 shows the grouping and fragility relationships suggested
by HAZUS.
22
Table 2.1 HAZUS grouping and fragility relationships for bridge classes in
California.
Year Description (acronym in Fragility values in terms of Sa-1.0 s
Class
built HAZUS) Slight Moderate Extensive Complete Dispersion
Major bridge
HWB1 < 1975 0.40 0.50 0.70 0.90 0.6
Length > 150 m
Major bridge
HWB2 1975 0.60 0.90 1.10 1.70 0.6
Length > 150 m
HWB3 < 1975 Single span 0.80 1.00 1.20 1.70 0.6
HWB4 1975 Single span 0.80 1.00 1.20 1.70 0.6
Multi-column bent,
HWB6 < 1975 0.30 0.50 0.60 0.90 0.6
Simple support, Concrete
Multi-column bent,
HWB7 1975 0.50 0.80 1.10 1.70 0.6
Simple support, Concrete
Single column, Box-girder,
HWB8 < 1975 035 0.45 0.55 0.80 0.6
Continuous concrete
Single column, Box-girder,
HWB9 1975 0.60 0.90 1.30 1.60 0.6
Continuous concrete
Continuous concrete
HWB10 < 1975 0.60 0.90 1.10 1.50 0.6
(not HWB8/ HWB9)
Continuous concrete
HWB11 1975 0.90 0.90 1.10 1.50 0.6
(not HWB8/HWB9)
Multi-column bent, Simple
HWB18 < 1975 0.30 0.50 0.60 0.90 0.6
support, Prestressed concrete
Multi-column bent, Simple
HWB19 1975 0.50 0.80 1.10 1.70 0.6
support, Prestressed concrete
Single-column, Box-girder,
HWB20 < 1975 Prestressed concrete 0.35 0.45 0.55 0.80 0.6
continuous
Single-column, Box-girder,
HWB21 1975 Prestressed concrete 0.60 0.90 1.30 1.60 0.6
continuous
Continuous concrete (not
HWB22 < 1975 0.60 0.90 1.10 1.50 0.6
HWB20/HWB21)
Continuous concrete (not
HWB23 1975 0.90 0.90 1.10 1.50 0.6
HWB20/HWB21)
HWB28 All other bridges that are not classified 0.80 1.00 1.20 1.70 0.6
The salient features noted from the critical review of HAZUS grouping and fragility
• HAZUS classifies bridges in two design eras, pre-1975 and post-1975. However,
1971 San Fernando and the 1989 Loma Prieta earthquakes. The extensive damage from
the 1989 Loma Prieta earthquake forced Caltrans to solicit the Applied Technology
23
Council (ATC) to conduct a detailed study and provide recommendations for design
A study by Ramanathan et al. (2015) showed that fragility curves are highly influenced
• HAZUS classifies the bridge classes that are not addressed in the main classification as
the bridge group. The other bridge group represents the high-risk bridge inventory.
This classification leads to a situation where multi-frame bridges that are not addressed
explicitly in the main group are in the non-classified group, although the seismic
vulnerability of slab bridges is much lower than the vulnerability of continuous box-
girder bridges (Mehr and Zaghi, 2016). Therefore, HAZUS classifications significantly
overestimate the seismic vulnerability and loss assessment for multi-frame bridges.
• Although HAZUS classifies bridges based on abutment type (monolithic versus non-
seismic notions), HAZUS does not suggest explicit fragility relationships based on
abutment type. Previous studies (Mangalathu et al., 2016a; Ramanathan et al., 2015)
have noted that the demand models and fragilities for various components and bridge
systems differ drastically depending on the abutment style. Further, Ramanathan et al.
(2015) indicated that diaphragm abutments are less vulnerable than seat abutments in
and simplified two dimensional analyses, and did not account for uncertainties in
geometric and material attributes for bridge classes such as the number of spans, span
24
length, deck width, and column height. Also, other researchers (Porter, 2010;
Ramanathan, 2012) have criticized the capacity spectrum method (CSM) of structural
analysis used in HAZUS. The capacity spectrum method estimates the capacity of
bridge in the form of a pushover curve of the column and the demand in the form of a
pointed out by Ramanathan (2012), columns are not always the critical components;
neglecting the damage to bearings, abutments, and shear keys underestimates the bridge
vulnerability.
• HAZUS suggests the same fragility relationships for bridge classes HWB10 and
HWB22, and for HWB11 and HWB23. It can be inferred from these grouped fragility
not a significant parameter for the bridge fragilities. This is consistent with similar
2016a). However, slab bridges, T-girder, and box-girder bridges are classified in the
same group and Ramanathan (2012) showed that these bridge classes do not have
and HWB11, HWB18 and HWB20, and HWB19 and HWB21 shows that single-
column bents (SCBs) are more vulnerable than multi-column bents (MCBs).
Ramanathan et al. (2012) and Mangalathu et al. (2016a) showed that SCBs are less
• While comparing the fragility relationships of bridge classes HWB22 and HWB23, and
HWB10 and HWB11 for moderate, extensive, and complete damage states, the effect
25
of design eras does not have an influence on the fragility relations. Such a conclusion
contradicts the research explicitly focusing on the effect of design eras (Mangalathu et
• HAZUS suggests the same fragility relationships for single-span bridges irrespective
of the design eras (HWB3 and HWB4). Although it might hold for bridges with
diaphragm abutments, it is clearly not the case for bridges with seat abutments as there
is an increase in the seat-width provision for newer era bridges. Since span-unseating
or bearing displacement is the critical component for single-span seat abutment bridges,
new era single-span seat abutment bridges are less vulnerable than their counterparts
vulnerable than continuous bridges. The study by Ranf et al. (2007) utilized damage
data collected from the Nisqually Earthquake in 2011 to reveal that this is not true for
lower damage states. As there is not enough data for higher damage states, it is not
certain whether the HAZUS fragility relationships (that is, that simply-supported
bridges are more vulnerable than continuous bridges) hold for higher damage states.
The study also indicated that HAZUS fragility relationships overestimate the damage
• Although HAZUS classifies bridges without considering the number of frames, Mehr
and Zaghi (2016) used three-dimensional nonlinear time history analysis to show that
single frame bridges do not have similar fragility curves to multi-frame bridges.
• An extensive plan review of the California bridge inventory revealed various column
major portion of concrete bridges in California and recent studies (Mangalathu et al.,
26
2016a; 2016b) have shown that bridges with circular and rectangular column cross-
• Although HAZUS classifies the bridges based on length (length > 150 m and length <
150 m), it is not clear whether the length is per frame or the total length of the bridge.
functionality in HAZUS and the damage state definitions used in the fragility analysis.
modification factor, three-span bridges are less vulnerable than their counterparts. A
Given the key points noted from the critical review of HAZUS bridge classification,
it is clear that HAZUS groupings and fragility relationships need significant improvement.
Also, it is rational to advance the grouping of bridge classes from a traditional perspective
As stated previously, the prevalent approach for the generation of fragility curves
is the convolution of demand models with capacity models. Demand models are usually
obtained by conducting non-linear time history analysis (NLTHA) on bridge models (Choi,
2002; Gardoni et al., 2003; Kim and Shinozuka, 2004; Mackie and Stojadinovic, 2001;
Mackie and Stojadinović, 2005; Banerjee and Shinozuka, 2007; Padgett, 2007;
27
Mangalathu et al., 2015; Ramanathan et al., 2015; Mangalathu et al., 2016a). In the
sampling across the uncertain input parameters (Padgett, 2007; Ramanathan et al., 2015,
Mangalathu et al., 2016a; Mangalathu et al., 2015a). It is highly likely that various
fragility curves are intended for the regional risk assessment of bridges (Mangalathu et al.,
2016a). The source of uncertainties can either be due to lack of knowledge (epistemic) or
Figure 2.5– Uncertainty sources for system demand and capacity (Ji et al., 2007).
Uncertainties can present both in seismic demand and capacity (Ji et al., 2007,
displacement without failure. Researchers have attempted to determine the extent to which
uncertainties affect the seismic demand, capacity, and fragilities. The uncertainties in
ground motions can be accounted for by including many records of ground motions to
cover as many frequencies and seismic energies as possible (Ji et al., 2007). However, the
28
number of records needed to have a reliable estimate of the fragility is not well defined
(Haselton et al., 2012). Celik and Ellingwood (2010) noted that uncertainties in ground
motion dominated overall uncertainty in structural response in the case of gravity load
designed reinforced concrete frames. Their study concluded that other sensitive parameters
affecting the seismic response of reinforced concrete frames are damping, concrete
strength, and joint cracking strain. The uncertainty in the capacity is usually accounted for
by modeling the capacity or limit state as a random variable (Ji. et al. 2007). For example,
HAZUS suggests a dispersion measure to account for variability in the damage state. In
the case of bridges, researchers have attempted to assess the sensitivity of seismic demand
or have evaluated fragility to the parameter uncertainty (Dicleli and Bruneau, 1995;
Nielson and DesRoches, 2006; Padgett, and DesRoches, 2007; Padgett et al., 2010; Ghosh
supported and continuous slab-on-girder steel bridges using linear elastic and nonlinear
inelastic analysis. Based on elastic spectral analysis, they noted that bearing forces, both in
longitudinal and transverse direction, were proportional to span length. The authors
concluded from the inelastic time history analysis of bridges that 3-lane bridges are less
vulnerable than 2-lane bridges. Another conclusion from their study is that the bridge
response was significantly influenced by the stiffness with which the steel bearings are
modeled. However, this conclusion was contradicted by the research on simply supported
steel-girder bridges conducted by Ala Saadeghvaziri and Rashidi (1998), in which the
bridge response was dependent on the stiffness of the bearings only in the transverse
Nielson and DesRoches (2006) carried out an experimental design to ascertain the
supported steel girder bridges in the central and southeastern United States (CSUS).
29
Nonlinearities in the abutments, bearings, columns, and bent caps were explicitly
considered in their study using detailed 3-D nonlinear models. The authors used a statistical
analysis called ANOVA to identify the significant parameters for bridge samples subjected
to seismic loading. The study revealed that damping ratio and loading direction are the
most important parameters affecting the seismic response of bridges. The study also noted
that column ductility and bearings deformations are sensitive to the type and stiffness of
bearings.
Padgett and DesRoches (2007) extended the procedure used by Nielson and
DesRoches (2006) to retrofitted bridges in CSUS and concluded that fragility curves
developed with sensitive parameters are nearly identical to those developed with all
potential sources treated as variables. Their study illustrated that preliminary screening of
parameters could reduce the simulation and computational efforts for the generation of
fragility curves. It has been noted from their study that the uncertainty in ground motion
attributed to the bridge modeling. The sensitivity study was further extended by Padgett et
al. (2010) to identify the effect of liquefiable soil and modeling parameters on the seismic
reliability of critical components of steel bridges in CSUS. Although such studies are
valuable for providing critical insights, most of them are rooted in rigorous statistical
analysis based on experimental design and demand exhaustive computational efforts. Also,
(2006) investigated the effect of foundation flexibility and soil-structure interaction on the
seismic demands. However, the demand model is conditioned only on one parameter (IM)
and hence it is difficult to estimate the sensitivity of other parameters on seismic demand.
Ghosh et al. (2013) used a multi-parameter demand model to account for the effect of all
uncertain parameters in the generation of fragility curves. The authors used four surrogate
30
modeling techniques, including polynomial response surface models, multivariate adaptive
regression splines, radial basis function networks, and support vector machines to
determine the best-fitting parameterized demand models involving the uncertain input
that all of the uncertain parameters have a significant influence on the seismic demand
model.
Kameshwar and Padgett (2014) identified the significant parameters that can affect
sequential forward selection scheme. The authors discussed the earthquake and hurricane
risks to multiple-span simply supported (MSSS) concrete girder bridges in South Carolina
Jeon et al. (2016) employed a Bayesian framework to screen the parameters that
California. The authors identified that parameters such as damping ratio, mass factor,
longitudinal gap, backfill type, pile stiffness, column longitudinal reinforcement ratio,
rotational and transverse stiffness of footing, bridge angle, main span length, side span-to-
main span length ratio, and column height have a significant influence on the seismic
demand of bridges. The study concluded that seven parameters (the earthquake direction
factor, concrete strength, rebar yield strength, coefficient of friction, and shear modulus of
elastomeric bearing pads, transverse gap, and abutment height) have little impact on all
expensive because they performed a set of stepwise regressions until the reduced model
31
demand and fragilities of bridge classes. The literature review on the uncertainty treatment
of input parameters in bridge fragilities suggests the sensitivity parameters vary depending
on the type of bridge and its location. Although these studies have been invaluable in
and fragilities of bridges, there is still a need for identifying the relative impact of each
uncertain input variable and the level of treatment needed for these variables in the
estimation of seismic demand models and fragility curves. As the current study focuses on
the box-girder bridges in California, a sensitivity study is needed for the box-girder bridges
in California that can (1) identify the variables that exhibit strongest influences on seismic
demand and seismic fragilities; (2) provide insight in quantifying whether the variation of
parameters which have a minimal influence on seismic demand and reduce unnecessary
and exhaustive efforts in statistical sampling; (4) identify parameters which could reduce
the uncertainty in demand models and fragility curves by more explicit evaluation of the
uncertainty distribution (e.g., by developing an extensive database); and (5) help bridge
(e.g. data collection, field investigations, censoring) on parameters that have significant
2.6 Closure
earthquakes. It has close to 29,000 bridges with varying ages and design parameters based
develop unique fragility curves for each structure across a regional portfolio. One strategy
that has been used to address this challenge is to group bridges into classes based on similar
design or structural performance. Traditionally, this grouping has been conducted based on
32
classification yields distinct seismic performances (or seismic demand) between the
grouped bridge classes. The literature review suggests the need for a performance-based
Fragility curves, which are probabilistic tools used to assess seismic damage to
highway bridges, can be generated based on expert opinion, empirical approach, and
through numeric methods. A review of the current methods for generating fragility curves
bridge fragility models developed in the 1990s to support loss estimation by the Federal
Emergency Management Agency. By necessity, these early models were derived with
simplified analysis methods, compared to a limited set of damage observations, and use a
bridge taxonomy based on the limited data fields available in the National Bridge Inventory
(NBI). HAZUS fragility relationships suffer major drawbacks which are discussed in detail
in this chapter. It has been noted that using HAZUS relationships leads to non-realistic
estimation of the seismic risk. Also, HAZUS framework is not well aligned with Caltrans
seismic design philosophy or the California bridge inventory. There is a need for the
generation of fragility curves that can lead to a realistic estimation of seismic risk in
California. High fidelity three-dimensional analytical models will be used in the current
The literature review also reveals the need for a sensitivity study on box-girder
bridges that can identify the significant input parameters and the relative impact of these
parameters on seismic demand models and fragilities. Such a sensitivity study would help
Caltrans to spend their resources (e.g. data collection, field investigations, censoring)
33
CHAPTER 3 MODELING OF BRIDGE COMPONENTS
This section presents the various bridge components and the adopted numerical
modeling strategies. Consistent with the previous work of Ramanathan et al. (2015),
bridges are classified into pre 1971 design era (Era 11, hereafter), 1971-1990 design era
(Era 22, hereafter), post 1990 design era (Era 33, hereafter) based on evolutions in the
The components can be primarily classified into as superstructure and substructure. The
super structure includes the girders, deck slab, and parapet, while the substructure consists
of abutments, footing, bents (beam and columns), bearings, and shear key.
A typical layout of the numerical modeling for a two-span bridge is shown in Figure
3.2. The bridge is modeled in three dimensions to capture responses in both the longitudinal
and transverse directions, as well as their interactions. The modeling is carried out with
the finite element package OpenSees (Mazzoni et al. 2006), incorporating geometric and
material nonlinearities. Rayleigh damping is used in the non-linear time history analysis.
34
The ground motions representative of seismic risk in California are used in this study. The
two horizontal components of the ground motions are assigned simultaneously to the
longitudinal and transverse direction of the bridge, and the orientation is assigned
randomly. The effects of vertical acceleration and spatially variable ground motions are
not considered in this study. The following section explains the numerical modeling of
3.1 Superstructure
elements since it is expected that the deck will remain elastic during earthquake loading.
Transverse deck elements are modeled using elastic beam-column elements (rigid and
massless) and are connected to the columns using rigid links to ensure the moment and
force transfer between members. Translational and rotational springs are added to the base
of the column to simulate the behavior of the footing. Zero length elements capturing the
response of the abutment back fill soil and bi-directional force (abutment piles or frictional
surface) are connected in parallel and are connected to the transverse deck elements in the
case of diaphragm abutments. The abutment pile or friction surface model is selected based
on the type of footing, whether the abutment is resting on piles or on a spread footing. In
the case of cantilever abutments, the wall stem flexure is connected in series with the bi-
directional force springs. Bearing pad elements and pounding elements are also modeled
The following section presents detailed modeling considerations for various bridge
components. In general, the bridge components are divided into superstructure and
substructure.
35
The superstructure (bridge deck) typically remains elastic during an earthquake.
The superstructure in this study is modeled using elastic beam-column elements and is
3.2 Substructure
California bridges have different pier types (single column bents (SCB), multi-
column bents (MCB), pier walls or pile shafts), footing types (spread footing, shaft pile, or
3.2.1 Bents
The bents are modeled using a combination of displacement based beam column elements
and rigid links to cause moment and force transfer between the members of the bent. The
36
finite element discretization of a single column and two-column bent is shown in Figure
3.3.
3.2.2 Columns
Columns are one of the most vulnerable components in the event of an earthquake; the
majority of bridge seismic failures in the past have been attributed to column failures.
Displacement-based beam column elements with fiber-defined cross-sections are used in
this study to model the columns (Figure 3.3). Fiber cross-sections have the distinct
advantage of specification of unique material properties for different locations across a
member’s cross-section. For instance, confined concrete is used to represent the concrete
behavior in the core section of the column, while unconfined concrete is used to represent
the unconfined cover concrete. The Chang and Mander (1994) model is used to define the
monotonic stress-strain curves of confined and unconfined concrete. Material models for
the concrete section are shown in Figure 3.4, and Figure 3.5 shows the effectively confined
core concrete area for the circular and rectangular section. Suppose, fc represents the
unconfined strength of concrete, the maximum concrete stress (fcc) and corresponding
strain (cc) can be calculated as (Mander et al. 1988):
37
f cc f c f l k e
fl
f c (1 k e )
fc
(3-1)
f c CF
fl
cc co (1 k 2 )
fc
where fl is the effective lateral confinement factor, ke and k2 are coefficients that are
functions of the concrete mix and lateral pressure, and co is the unconfined strain in
concrete.
The effective lateral confining pressure (fl, Equation 3.2) for a circular column can
be calculated as:
1 4 Asp
fl s f yh ; s (3-2)
2 dss
where Asp is the area of transverse reinforcement bar, fyh is the yield strength of the
transverse reinforcement, s is the center to center spacing or pitch of spiral or circular hoop,
and ds is the diameter of spiral bar centers. This is shown in Figure 3.5. The ratio of the
38
area of the effectively confined concrete core to the area of core of section enclosed by the
perimeter lines of the perimeter spiral or hoop (Equation 3.3), ke can be calculated as
s'
1
2d s
ke (3.3)
1 cc
where cc is the ratio of area of longitudinal reinforcement to the area of core of section. k2
is given as 5k1.
ds ds
cover concrete cover concrete ds-s'/2
(spalls off)
A A
ineffectively s' s
confined core
45˚
effectively B B
¼s'
confined core
Section B-B Section A-A
a)
bc bc
effectively w'
confined ineffectively bc-s'/2
core confined
dc
core
Y dc-s'/2 Y s' s
cover Z Z
concrete
Section Z-Z (spalls off) Section Y-Y
b)
Figure 3.5 – Effectively confined core for a) circular hoop reinforcement and, b)
rectangular hoop reinforcement (Mander et al. 1988).
For a rectangular bc dc (Figure 3.5), the area of effectively confined concrete core to the
39
n
(w' ) 2 s' s'
1 i 1 1
ke
i 1 6bc d c 2bc 2d c
(3.4)
1 cc
The lateral confining stress of concrete in x and y directions (Figure 3.5) is given
as
Asx
f l x x f yh ; x
dc s
(3.5)
Asy
f l y y f yh ; y
bc s
where Asx and Asy are the total area of transverse bars running in the x and y direction.
The Menegotto and Pinto (1973) model, later modified by Filippou et al. (1983), is
Various cross sections such as circular, rectangular, wide perimeter, and wide
circular are noted from the plan review. The typical details of some column cross-sections
are shown in Figure 3.6. The wide perimeter cross-section corresponds to the section in
which the reinforcement is laid around the perimeter, while the wide circular cross-section
concrete core.
40
Figure 3.6 – Typical cross-sections noted from the bridge plan review (Caltrans,
2017).
Figure 3.7 represents the common choice for columns with circular distribution of
longitudinal reinforcement contained within transverse spirals or hoops. The area outside
the core concrete can be circular, hexagonal, octagonal, or any other shape. In the case of
a circular cross-section, flexural strength, shear strength, and moment capacity are
41
3.2.2.2 Rectangular columns
Rectangular columns of various sizes are noted from the plan review of bridges.
Figure 3.8 shows some of the typical rectangular cross sections noted from the plan review.
It is noteworthy to mention that the strength and stiffness of the rectangular columns
The superior performance of the spiral reinforcement over the rectilinear tie
transverse reinforcement led to the development of columns with interlocking spirals (also
called oblong columns, Figure 3.9). The confinement factor of a single column fiber is
confined by a single spiral and the interlocking region is a relatively small area (Correal et
al. 2007).
42
3.2.3 Idealization of bridge columns
some idealizations are used in this study. The idealization is carried out in such a way as
to mimic the bridge inventory noted from plan review and is detailed in Table 3.1. The
idealization of the cross-sections is carried out based on the area of confined concrete, as
the core concrete will continue to carry stress at higher strains. The cover concrete becomes
ineffective once the compressive strength is attained in the flexural deformation, and the
43
Table 3.1 – Idealization of bridge columns.
Column cross-section from the bridge inventory Idealized cross-section shape Remarks*
No idealization
needed
Do = Di
CF= circular
Idealized to
circular cross-
section. Area of
Symmetric
core concrete
remains the same
Dco = Dci
co = c i
CF= circular
No idealization
needed
Loo = Lco
Boo = Bco
CF= circular
44
Idealized to oblong
cross-section. Area
of core concrete
remains the same
Loc = Loi
Boc = Boi
co = ci
CF= circular
Idealized to
rectangular cross-
section.
L = Lr
B = Br
CF= rectangular
Idealized to
Wide perimeter
rectangular cross-
section.
Ll = Lc
Bl = Bc
CF= rectangular
Idealized to
rectangular cross-
section. (Area of
two sections
remains the same)
Blc = Brc
CF= rectangular
45
Idealized to
rectangular cross-
section. (Area of
two sections
remains the same)
Brcc = Bric
CF= rectangular
Idealized to
rectangular cross-
section. (Area of
two sections
remains the same)
Bro= Bio
CF= rectangular
Planning to
exclude as the
Wide circular
46
3.2.4 Validation of bridge columns
Figure 3.10 shows the comparison of the numerical model with the experimental results
for Era 33 columns. The geometric and reinforcement details are obtained from Lehman
and Moehle (2000), and the modeling details are outlined in section 3.2.2. It is seen from
the comparison that the numerical model is able to capture the key responses fairly well.
a) b)
b)
Figure 3.10 – Comparison of experimental and numerical results for era 33
columns (Lehman and Moehle, 2000) a) Specimen No. 415 b) Specimen No. 815.
3.3 Abutments
Abutments can be classified in two basic types: diaphragm abutments and seat
abutments (Ramanathan, 2012). Diaphragm abutments are cast monolithic with the
superstructure. As they engage the backfill soil during seismic action, diaphragm
abutments provide a good source of energy dissipation and reduce the likelihood of span
restrained longitudinally by the abutment backwall and transversely by the piles and the
shear key. The stiffness and resistance to the seismic action increases when the deck is in
contact with the abutment backwall in the longitudinal direction. However, as the
47
superstructure moves away from the abutment, the resistance depends primarily on the
bearing pads, which makes it susceptible to unseating. The backwall of the seat abutment
is typically designed to fail under impact and passive response, before damaging forces are
various design eras are shown in Figure 3.11. Abutment can be on piles, on spread footing
or cantilever. Abutments resting on piles are the major configuration for both diaphragm
and seat abutments. However, the review of bridge plans also revealed a variety of unusual
abutment wall details (Figure 3.12) which seemed to be designed to provide a weak link in
the stem wall above the footing which could either translate along a construction joint or
rotate on an intermediate section of the stem wall. These details were found on about 15-
48
Figure 3.12 – Unusual abutment types (Caltrans, 2017).
structural response. The earth pressure on the abutment is due to the longitudinal response
of the bridge deck and includes passive and active resistance. Passive resistance is provided
by the backfill soil and piles/friction surface (depending on the abutment footing type); it
develops when the abutment moves toward the backfill soil. Piles/friction surface alone
contribute the active resistance, which is activated when the abutment moves away from
the backwall soil. The passive response of the abutment backwall is simulated using the
hyperbolic soil model proposed by Shamsabadi and Yan (2008) and is given in Figure 3.13.
University of California Los Angeles with 5.5 ft. high backwalls and typical non-cohesive
and cohesive backfill soils. The test results were then extended to develop closed form
solutions for the abutment backfill soil response for a range of backwall heights based on
logarithmic-spiral failure surfaces coupled with a modified hyperbolic soil stress strain
behaviour. Fult is the maximum abutment force developed at maximum displacement ult.
49
Equation 3.6 presents the closed form solution for the force displacement response of the
backfill soil, where F is the force expressed in kip/ft width of the backwall, is the
displacement expressed in inches, and H is the height of the backwall expressed in feet.
8
F () H1.5 Granular backfills
1 3
(3.6)
8
H Cohesive backfills
1 1.3
inches) for granular (sandy soils) and cohesive (clay soils) backfills, respectively, and
substitution of these values in Equation 3.6 yields the ultimate force in the abutment. The
which corresponds to the model proposed by Shamsabadi and Yan (2008). The force-
F () (3.7)
1
Rf
K max Fult
where, the failure ratio Rf is 0.7, Kmax is secant stiffness at half of the maximum abutment
force.
50
Figure 3.13 – Modeling of the abutments.
Piles provide longitudinal and transverse stiffness to the abutments when abutments
rest on piles. The trilinear force-deformation response of the pile, along with the associated
modeling parameters, are presented in Figure 3.13. The initial yield parameters (1, F1) are
determined following the design recommendations of the Caltrans 2014 draft of bridge
design aids on ‘Permissible Horizontal Loads for Standard Plan and Steel HP Piles’
(Caltrans, 2015). The plastic yielding parameters (2, F2) are calculated based on results of
modeling various pile systems simulated in LPILE (Caltrans, 2015). The hysteretic
behavior of piles is captured using the Hysteretic material in OpenSees with the hysteretic
parameters pinchX and pinchY as 0.75 and 0.5 (Ramanathan, 2012). In contrast, for
abutments supported on spread footings, a frictional response model is used. The maximum
force (Fs) is calculated as the product of the coefficient of friction (f) and the dead load
reaction on the abutment. For cantilever abutments, the bi-linear model is obtained from
51
3.4 Bearings
The bearings most commonly used for seat abutment bridges are rocker or
different response mechanisms. Rocker bearings shown in Figure 3.14 (a) are considered
vulnerable due to non-ductile transverse keeper plate failure and longitudinal instability
(Mander et al., 1996). The elastomeric bearings shown in Figure 3.14(b) usually transfer
horizontal forces using friction, and their behavior is characterized by sliding. Elastomeric
bearings decouple the superstructure from the substructure, and thus the superstructure is
bearing can also be found in Era 11 bridges. The predominant difference between the Era
11 and the later design eras is the replacement of steel rocker bearings with elastomeric
steel bearings. This shift in the bearing type is due to advancement in the seismic design
philosophies.
(a) (b)
Figure 3.14 – Types of bearings: (a) rocker bearing and (b) elastomeric bearing
(Mangalathu et al. 2016).
The elastomeric bearing is assumed to be elasto–plastic and the yield force, Fy, is
obtained by multiplying the normal force by the coefficient of friction (Figure 3.15). The
52
rocker bearing is modeled following the experimental work of Mander et al. (1996). This
behavior is shown in Figure 3.15. An elasto–plastic behavior is assumed for the friction
bearings. The bearings are modeled using the Steel01 material in OpenSees.
Shear keys help restrain the relative transverse movement between the deck and the
bridge abutments. A shear key can fail through four mechanisms in the event of an
earthquake, namely shear friction, flexure, shear, and bearing (Megally et al. 2002). The
53
(conventional designs) (Caltrans 2017). Since the isolated shear key is a new type of design
and does not exist in the current inventory, this study will focus only on the non-isolated
shear keys. The nonlinear model of the shear key is also depicted in Figure 3.16. Fcap
denotes the capacity of the shear key, which is computed as the product of the dead-load
reaction and the acceleration (Caltrans 2015). Megally et al. (2002) conducted a series of
experiments on the shear keys and found that Δmax minus Δgap equal to 3.5 in. is the
deformation at which the capacity of the shear keys essentially degrades to zero.
3.6 Pounding
Seismic pounding is the impact between the bridge decks, between deck and
direction. Impact occurs when the relative displacement between adjacent decks or deck
and abutment exceeds the gap between them. Significant pounding damage was noticed at
the I-5/SR-14 interchange during the 1994 Northridge earthquake (Muthukumar and
Figure 3.17 – Pounding damage in bridges during the 1994 Northridge earthquake:
(a) barrier rail damage and, (b) connector collapse (Muthukumar and DesRoches,
2006).
model the pounding between superstructures and abutments. This material model explicitly
54
accounts for the loss of hysteretic energy (Figure 3.18). The maximum deformation, m, is
assumed to be 1.0 in. The yield deformation, 1, is assumed to be 0.10m. The stiffnesses,
k1 and k2, are recommended to be 1022.3 kip/in/ft and 351.755 kip/in/ft, respectively.
Figure 3.18 – Analytical Model for pounding between deck and abutment back wall.
3.7 Foundation
The foundation provides a means to transmit service and ultimate loads from the
structure to the underlying soil. Foundations can be classified as either shallow or deep. As
the name implies, the loads from the structure are transferred to the underlying soil at a
shallow depth for shallow foundations. Deep foundations are used when soil conditions are
not favourable to shallow foundations and transfer the load through piles. The type of
foundation for a particular bridge is determined by various factors such as soil conditions,
overhead clearance, existing utilities, and proximity to existing facilities such as buildings
and railroads (Caltrans 2017). The possible types of footings are shown in Figure 3.19.
Foundations are modeled using elastic translational and rotational springs (Figure 3.20)
55
Figure 3.19 – Bridge foundation types (Priestley et al. 1996).
56
Figure 3.20 – Modelling of foundations.
3.8 Closure
various component models, to generate a global analytical model of the bridge for fragility
analysis. Displacement-based beam column elements are used to model the columns.
Translational and rotational springs are added to the base of the column to simulate the
behavior of the footing. Zero length elements capturing the response of the abutment back
fill soil and bi-directional force (abutment piles or frictional surface) are connected in
parallel and to the transverse deck elements in the case of diaphragm abutments. Bearing
pad elements and pounding elements are also modeled with zero length spring elements
and are connected in parallel to abutment springs in the case of seat abutments.
Interested readers are directed to the study from our research group members
(Ramanathan et al. 2015; Soleimani 2017), for the validation of the global analytical model
of the bridges. Ramanathan et al. (2015) compared the efficacy of the nonlinear dynamic
analysis of the bridge models with the recorded ground motion data. The authors compared
the response of a two span reinforced concrete box-girder bridge built in 1971 (Meloland
Road Overpass) with the recorded ground motion during the 1970 Imperial Valley
57
peak ground acceleration (PGA) values of the ground motions that strike the bridges were
of intensity, 0.32 g, 0.30 g, and 0.23 g in the longitudinal, transverse, and vertical
directions, respectively. The authors concluded that the analytical model yield comparable
results to the sensor data for the Meloland bridge. Note that the lack of data prevents the
comparison of the bridge model with the ground motions that can cause significant
nonlinear response. Soleimani (2017) compared the dynamic response of bridge columns
with the full scale shake table test on single column bridge bent (Schoettler et al. 2012) to
the various ground motions. It is noted from their study that the analytical model can predict
the shear force and deformation of the column fairly well for various damages states.
58
CHAPTER 4 PERFORMANCE BASED GROUPING OF BRIDGE
CLASSES
infrastructure systems is the generation of fragility curves that are applicable to a class of
bridges. California, a state with a high seismic hazard and a history of damaging
earthquakes, has close to 29,000 bridges that vary in age based on their construction.
However, it is cumbersome and time-consuming to develop unique fragility curves for each
structure across a regional portfolio. One strategy that has been adopted to address this
challenge is to group bridges into classes with similar design or structural performance.
The bridges in a particular class (or group) are expected to have similar performance or
design attributes that yield distinct seismic performance to bridges is an important step in
this procedure.
bridge classes with unwarranted grouping (HAZUS, 2013). The limitation of the HAZUS
grouping is discussed in detail in the literature review of this thesis. This chapter explains
the various performance based grouping strategies such as analysis of variance (ANOVA),
approaches are used widely in many disciplines such as biology, medical, and industrial
engineering (Vidakovic 2011; Vickers 2005), the application, relevance, and advantages
of these grouping techniques for the grouping of bridge classes for probabilistic seismic
performance assessment have not been fully explored. The selected methods have different
approaches and underlying assumptions to grouping structures, which are reviewed and
compared on the basis of statistical power to identify attributes that dictate distinct bridge
59
sub-classes of structural performance. The comparison is based on case studies of two-span
and three-span bridges in California. The assumptions underlying each approach are
approach is suggested in this chapter to group the California bridge inventory. A review of
the application of ANOVA, ANCOVA, and KW in grouping the bridge classes is given in
This section briefly introduces the ANOVA, ANCOVA, and KW methods for
various bridge sub-classes and by identifying whether they are statistically different. The
response considered in this study includes (1) maximum column curvature ductility, (2)
maximum abutment passive deformation, (3) maximum abutment active deformation, (4)
maximum abutment transverse deformation, (5) maximum unseating deformation, and (6)
maximum bearing deformation. The maximum responses are obtained from a set of
NLTHAs. Figure 4.1 shows the scatter plot of seismic demand or response (D) of two
typical bridge groups with the IM, along with the probability distribution of their seismic
demands. PSDMs are usually obtained by performing a linear regression for a pair of D
and IM, using a suite of N ground motions. In a mathematical form, PSDM can be written
as
where a and b are the regression coefficients and, Sd is the median estimate of the demand
in terms of an IM. The coefficients a and b are obtained by performing a linear regression
60
N
ln d ln ( S d )
2
i
d / IM i 1
(4.2)
N 2
are any significant differences between the means of two or more independent, unrelated
groups (Miller Jr 1997; Keselman et al., 1998; Vidakovic 2011; Mangalathu et al., 2017a;
Mangaalthu et al., 2017b). It tests the hypothesis ( H 0 : D1 D2 ,..., Dk ) that the mean seismic
responses of different bridge classes are equal (Figure 4.2) and can group the bridge classes
accordingly. The assumptions underlying ANOVA are: (1) the responses are mutually
independent, (2) homogeneity of response variance, and (3) samples are mutually
independent. If the assumptions are violated, ANOVA is not a powerful test and the results
Let Di1, Di2,…,Din, be the samples of bridge responses and n be the sample size,
then N i 1 ni is the total sample size and k the number of groups to be compared.
k
61
H 0 : D1 D2 ... Dk
(4.3)
H1 ( H 0 ) c (or i j for atleast one pair i, j )
The total sum of squares (SST) is represented as a sum of the treatment sum of
SSTr / (k 1)
F (4.5)
SSE / ( N k )
with k-1 and N – k degrees of freedom. The null hypothesis is rejected if the F - statistic is
The p–value is the evidence against a null hypothesis or the probability that the
variation between groups occurred by chance. The p–value can be interpreted as the
Hypothesis
H0: D1 = D 2 ,…,= D k
Frequency
Ha: D 1 ≠ D 2 ≠,…, D k
… n
2
62
4.3 ANCOVA based grouping
formulation includes the bridge response, bridge sub-class, and ground motion intensity
relationship between the bridge response and the intensity measure for different bridge sub-
classes (or PSDMs) and to identify whether slopes of the regression lines are significantly
different. If the slopes are not significantly different, a regression line is drawn through
each group of points with the same slope. The intercept of the regression lines is then
checked, and if they are different, it can be concluded that the response of the bridge sub-
classes is different (Figure 4.1). The statistical significance of treatment is more likely to
where ln(Sd) is the response variable, ln(IM) the overall population mean of the intensity
1
measure, at the number of treatments, n the common sample size, ln( IM ) ln( IM )ij
at n i , j
the
overall mean of ln(IM)’s, bs the regression slope, and si the treatment effect. The error
term ij is assumed to be an independent normal distribution with zero mean and variance
2. ln( IM )i
1
ln( IM )ij is the ith bridge attribute effect for the ln(IM)’s. The means ln( S d )
n j
and ln( S d )i are defined analogously to ln( IM ) and ln( IM ) i , respectively. The sum of
63
at n at n
Sln( IM ) ln( IM ) (ln( IM )ij ln( IM )) 2 , Sln( IM ) ln( Sd ) (ln( IM ) ij ln( IM ))(ln( S d ) ij ln( S d ))
i 1 j 1 i 1 j 1
at n at
Sln( Sd ) ln( Sd ) (ln( S d )ij ln( S d )) 2 , Tln( IM ) ln( IM ) (ln( IM ) i ln( IM )) 2
i 1 j 1 i 1
at at
Tln( IM ) ln( Sd ) (ln( IM )i ln( IM ))(ln( Sd )i ln( Sd )), Tln( S
i 1
d ) ln ( Sd )
(ln( S d ) i ln( S d )) 2
i 1
(4.8)
at n at n
Qln( IM ) ln( IM ) (ln( IM )ij ln( IM )i ) 2 , Qln( IM ) ln( Sd ) (ln( IM ) ij ln( IM ) i )(ln( S d ) ij ln( S d ) i ),
i 1 j 1 i 1 j 1
at n
Qln( Sd )ln ( Sd ) (ln( S d )ij ln( S d )i ) 2
i 1 j 1
2
SSE Q ln( IM ) ln( Sd )
s 2 MSE , where SSE Qln( Sd ) ln( Sd ) (4.9)
(at (n 1) 1) Qln( IM ) ln( IM )
S 2ln( IM ) ln( Sd )
SSE ' Sln( Sd ) ln( Sd ) , with at n 2 degrees of freedom (4.11)
Sln( IM ) ln( IM )
The p–value of the F–statistic (upper trail area of the F–distribution) can be calculated as
64
where Fcdf is the cumulative distribution function of the F–distribution. The p–value is the
Equation 4.7 is zero. The inclusion of ground motion intensity in the grouping of bridge
sub-classes has several advantages: there is a higher probability that the test will reject a
false H0; there is a reduction in bias caused by chance differences between groups; and
Huietema 1980). ANCOVA assumes the homogeneity of regression slopes; and if the
reduction in the statistical power (Owen and Froman, 1998). This limitation can be averted
The stringent assumptions on the response data placed in ANOVA and ANCOVA
are often difficult to satisfy, which has led the development of non–parametric and semi–
parametric version of ANOVA. KW tests place no restriction on the population data, and
are applicable to complex experiments and messy sampling plans. The KW test statistic
(H) for k bridge sub-classes with sample sizes, n1,…,nk is (Kruskal and Wallis, 1952)
1 k Ri 2 ( N 1) 2
H' N (4.15)
S 2 i 1 ni 4
where, N ni , and Ri is the sum of the ranks for the ith sample, and
1 k ni
( N 1) 2
S2 R N
2
(4.16)
N 1 i 1
ij
j 1 4
65
where Ri is the sum of the ranks for the ith sample across the sample size j. If there are no
ties in the data, Equation 4.15 can be simplified to
ni ( N 1)
2
12 k
1
H
N ( N 1) i 1 ni Ri
2
(4.17)
hypothesis that responses from different bridge sub-classes have identical distribution
functions against the alternate hypothesis that the samples differ only with respect to the
median, if at all. Despite all its advantages over ANOVA and ANCOVA, the power of the
Two-span and three-span box girders bridges, which possess a major portion of the
California bridge inventory (Ramanathan, 2012) are the subject bridges in this study. The
selected bridges were designed and constructed prior to 1970. A typical layout of a two-
span box-girder bridge is shown in Figure 4.3. The current study adopts six different bridge
attributes such as (1) bearing type (elastomeric or rocker bearings), (2) column cross–
spread footing), (4) abutment backfill (clay or sand), (5) interior bent type (single–column
concrete).
66
Figure 4.3 – General layout of a two–span concrete box–girder bridge.
bridges, and are given in Table 4.1, which shows the mean value (), standard deviation
(), and the associated probability distribution of various input variables. These variables
are determined based on an extensive plan review of California bridges. In addition to the
uncertainties related to structures, the uncertainty in ground motions are accounted for
using the suite of ground motions assembled by Baker et al. (2011). The entire suite of
ground motions are scaled by a factor of two (Ramanathan, 2012) to have sufficient
67
response data of IMs higher than Palmdale spectrum (the highest probabilistic design
hazard level in California), and thus the expanded suite of 320 ground motions is used for
the current study. Additionally, the spectral acceleration at 1.0 sec (Sa–1.0s) is the optimal
intensity measure for the class of concrete box–girder bridges (Ramanathan, 2012), and is
As mentioned before, the objective of this study is to identify whether all of the
grouping techniques lead to similar sub-classes. To compare the responses with different
bridge attributes, the total simulation is split amongst the bridge attributes. For example, to
(160) are carried out for bridges with rectangular columns, while the remaining simulations
are performed for bridges with circular columns. The response variable and the ground
motion intensity measure are transformed into the lognormal space to produce a linear
relationship between the two (Cornell et al. 2002; Managalathu et al. 2016b). Using the
relationship between the transformed response variable and intensity measure, ANOVA,
ANCOVA, and KW grouping techniques are carried out. In all the grouping methods, p–
values are computed to interpret the results of the hypothesis test. A smaller p–value
indicates stronger evidence for rejecting the null hypothesis (H0), and a cut–off p–value of
0.05 (Mangalathu et al. 2016a) is adopted in the current study. For example, at a cut–off
value of 0.05, if the p–value is less than 0.05, it can be concluded with a 95% degree of
confidence that the variation in the demand measure is not due to random chance, but due
to the influence of the different bridge attributes. p–values less than 0.05 are highlighted in
Table 4.2. The table indicates that the various demand parameters are typically sensitive to
different bridge attributes. Among six bridge attributes, the type of interior bent is a
significant parameter for all cases. It can be explained that the transverse moment demands
in single column bents are higher than the longitudinal moment demands and have less
68
The grouping pattern identified by the three methods is the same for 80% of the
cases. ANOVA and KW yield the same identification of significant bridge attributes in
96% of the cases. The results of ANCOVA are different from those of ANOVA and KW
in comparison to ANOVA and KW tests, which is shown in italics in Table 4.2. While
comparing the ANOVA and KW grouping methods, ANOVA identifies more relevant
parameters. The reason for such a discrepancy is investigated in detail in the following
section.
69
Table 4.1 – Uncertainty distribution considered in the bridge models
Distribution
Parameter Units
Type μ σ
Concrete compressive strength (fc) MPa Normal 29.03 3.59
Reinforcing steel yield strength (fy) MPa Lognormal 465.0 37.30
Span length (L)
Two-span mm Lognormal 31775 8738
Approach to main span ratio (three-span bridge) – Normal 0.57 0.13
Deck width (Bd)
Single column bent mm Lognormal 9780 1980
Multi-column bent mm Lognormal 11970 2418
Column height (H) mm Lognormal 6625 865
Abutment backwall height (Ha)
Diaphragm abutments
On piles mm Lognormal 3234 488
On spread footings mm Lognormal 2925 1056
Seat-type abutments
On piles mm Lognormal 2186 441
On spread footings mm Lognormal 2186 441
Abutments on piles - Lateral capacity/deck width (Kpa)
Diaphragm abutment N/mm Lognormal 1120 404
Seat-type abutment N/mm Lognormal 1498 540
Abutments on spread footing
Coefficient of friction (as) – Normal 0.40 0.075
Yield displacement (y) mm Uniform 19.0 13.4
Elastomeric bearing pad
Stiffness per deck width (Kb) N/mm/m Lognormal 908 327
Coefficient of friction for bearing pad (b) – Normal 0.30 0.10
Rocker bearing
Coefficient of friction (l, longitudinal direction) – Normal 0.04 0.01
Coefficient of friction (t, transverse direction) – Normal 0.10 0.02
Gap (g)
Longitudinal (btw. deck and abutment wall) mm Lognormal 23.5 12.5
Transverse (btw. deck and shear key) mm Lognormal 12.8 2.58
Mass factor (m) Uniform 1.25 0.007
Damping () Normal 0.045 0.0125
Acceleration for shear key capacity (as) g Lognormal 1.00 0.20
Longitudinal reinforcement ratio () (%) Uniform 2.25 0.52
Pile group – pile cap and piles
Translational stiffness (Kft)
Single column – 1% long. rebar N/mm Normal 297716 140101
Single column – 3% long. rebar N/mm Normal 245178 105076
Multi column – 1.5% long. rebar N/mm Normal 140101 105076
Rotational stiffness (Kfr)
Single column – 1% long. rebar N-m/rad Normal 4.5109 1.1109
Single column – 3% long. rebar N-m/rad Normal 6.8109 1.1109
70
Table 4.2 – p–values by ANCOVA, ANOVA and KW test
Column (μϕ) Passive (δp) Active (δa) Transverse (δt) Unseating (δs) Bearing (δb)
Ty
Bridge attribute ANC ANO ANCO ANO ANCO ANO ANCO ANO ANCO ANO ANCO ANO
pe KW KW KW KW KW KW
OVA VA VA VA VA VA VA VA VA VA VA VA
Column cross–
Diaphragm abutment
0.000 0.000 0.000 0.834 0.667 0.567 0.897 0.720 0.631 0.908 0.723 0.686 – – – – – –
section
Interior bent 0.000 0.000 0.000 0.038 0.233 0.326 0.046 0.315 0.417 0.000 0.000 0.000 – – – – – –
Abutment 0.000 0.000 0.000 0.494 0.354 0.322 0.382 0.359 0.332 0.000 0.021 0.033 – – – – – –
– – – – – –
Two–span bridges
Backfill 0.458 0.635 0.856 0.000 0.000 0.000 0.000 0.000 0.000 0.018 0.209 0.387
Superstructure 0.438 0.775 0.875 0.715 0.464 0.639 0.832 0.467 0.680 0.703 0.230 0.269 – – – – – –
Column cross–
0.043 0.317 0.343 0.149 0.446 0.554 0.463 0.727 0.800 0.005 0.048 0.064 0.263 0.466 0.509 0.116 0.223 0.318
section
Seat–type abutment
Bearing 0.005 0.003 0.002 0.043 0.008 0.001 0.124 0.031 0.003 0.008 0.002 0.000 0.000 0.000 0.000 0.002 0.000 0.008
Interior bent 0.000 0.000 0.000 0.001 0.000 0.000 0.000 0.000 0.000 0.007 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
Abutment 0.302 0.988 0.943 0.252 0.052 0.038 0.469 0.125 0.085 0.116 0.507 0.548 0.600 0.200 0.209 0.194 0.494 0.315
Backfill 0.045 0.032 0.012 0.148 0.401 0.478 0.145 0.486 0.537 0.549 0.892 0.831 0.722 0.844 0.887 0.641 0.667 0.747
Superstructure 0.631 0.413 0.321 0.039 0.609 0.686 0.084 0.583 0.667 0.036 0.643 0.646 0.001 0.438 0.451 0.351 0.600 0.600
Column cross–
0.010 0.135 0.070 0.690 0.882 0.873 0.709 0.897 0.859 0.426 0.405 0.470 – – – – – –
section
Diaphragm
abutment
Interior bent 0.000 0.000 0.000 0.000 0.093 0.151 0.000 0.077 0.128 0.000 0.000 0.000 – – – – – –
Abutment 0.095 0.520 0.356 0.317 0.298 0.529 0.440 0.368 0.620 0.441 0.954 0.953 – – – – – –
Three–span bridges
Backfill 0.799 0.637 0.771 0.000 0.002 0.002 0.000 0.007 0.005 0.228 0.568 0.626 – – – – – –
Superstructure 0.155 0.347 0.498 0.419 0.850 0.807 0.314 0.763 0.723 0.877 0.593 0.551 – – – – – –
Column cross–
0.097 0.451 0.459 0.582 0.548 0.625 0.858 0.853 0.881 0.001 0.027 0.025 0.637 0.907 0.935 0.669 0.569 0.914
Seat–type abutment
section
Bearing 0.258 0.378 0.482 0.005 0.043 0.357 0.000 0.002 0.048 0.006 0.066 0.173 0.000 0.000 0.003 0.128 0.168 0.754
Interior bent 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
Abutment 0.003 0.104 0.065 0.002 0.334 0.739 0.000 0.153 0.376 0.000 0.174 0.158 0.000 0.000 0.000 0.014 0.050 0.034
Backfill 0.000 0.001 0.001 0.388 0.565 0.952 0.918 0.958 0.836 0.142 0.272 0.259 0.001 0.059 0.072 0.023 0.117 0.058
Superstructure 0.172 0.961 0.678 0.088 0.776 0.815 0.067 0.885 0.994 0.005 0.321 0.327 0.004 0.718 0.779 0.083 0.732 0.931
71
4.6 Comparison of various grouping techniques
ANCOVA shows that some of the design attributes are significant, while ANOVA and KW
tests identify the same attributes as non–significant, as noted in Table 4.2. This non–significance
can be interpreted in two ways. Firstly, the lack of attribute significance may be obtained because
there is indeed very little or no true effect of the design attribute on the seismic response.
Alternatively, the finding suggesting a lack of significance can be attributed to the grouping
technique having less power to identify the effects of the design attribute (Huitema, 1980). To
evaluate the power and efficiency of various grouping techniques, the current study selects three
bridge sub–classes: (1) three-span diaphragm abutment bridges grouped by the cross–section shape
based on the curvature ductility response, (2) two-span seat-type abutment bridges grouped by the
cross–section shape based on the transverse abutment response, and (3) two-span seat-type
abutment bridges grouped by the abutment type based on the passive abutment response. The three
sub classes are carefully chosen such that one of the grouping techniques identifies the design
attribute as non–significant.
For the three-span diaphragm abutment bridge, two sub–classes are formed based on the
curvature ductility response according to the column cross–section: (1) with circular cross–section
(hereafter, 3SDC) and (2) with rectangular cross–section (hereafter, 3SDR). The ANCOVA result
(p–value of 0.010) reveals that the cross–section is significant for this bridge, while the ANOVA
(p–value of 0.135) and KW (p–value of 0.070) tests conclude it is not significant. Figure 4.4 shows
the data analysis results of the curvature ductility for 3SDC and 3SDR. The power of ANOVA and
the ANCOVA depends on the three or four assumptions and is evaluated initially. One of the
critical assumptions in both ANOVA and ANCOVA is that the data is normally distributed. The
normality of 3SDC and 3SRC is checked using the Kolmogorov–Smirnov test (KS test)
(Kolmogorov, 1933), which identifies the null hypothesis that the data is normally distributed. If
72
the p–value is less than = 0.05, the null hypothesis is rejected. Thus, there is enough evidence
that the data do not follow a normally distributed population. The p–value of normality check for
3SDC and 3SDR is 0.067 and 0.099, respectively, and hence one fails to reject the null hypothesis.
The assumption that each observation is mutually independent is valid because of the random
sampling and pairing of the bridge and ground motions. The validity of the assumption of the
independence of the treatment effect lies in the fact that the different bridge attributes are
independent of each other. ANCOVA has one additional assumption regarding the homogeneity
of regression slopes. As the demand measure increases with the intensity of ground motions, it
Figure 4.4 – Histograms and descriptive statistics for case 1: a) 3SDC, b) 3SDR, c) box plot
of 3SDC and 3SDR, and d) ANCOVA regression lines of 3SDC and 3SDR.
The homogeneity of regression slopes is checked in the current study using the F-test for
the equality of slopes (Vidakovic, 2005) and also seems to hold. ANCOVA and ANOVA have a
73
greater power than the KW test when the above assumptions are satisfied (Vidakovic, 2005;
Vickers 2005).
Figure 4.4 (c) shows box plots of the curvature ductility for 3SDC and 3SDR. The mean
value of intensity measure in the logarithmic scale for 3SDC and 3SDR is –1.14 and –1.21,
respectively. The ground motions are randomly assigned in this study because it is computationally
intensive and almost impractical to produce responses for all the cases with the same suite of
ground motions. The experiment has been designed in a way that the design attributes are randomly
sampled in the 320 simulations from a practical perspective. Such an assignment also helps assess
the combined effect of two or more different design attributes; for example, the combined effect
of the abutment type and column cross-section. The question arising from this case is whether the
column response of 3SDC and 3SDR can be compared as there is a clear variation in the IM and
the column response is highly related to IM. It is possible that the results are skewed if the column
responses are compared as the mean values of IM are different. ANOVA and KW tests neglect
such variation in IM and such a comparison leads to erroneous results. The mean value of the
curvature ductility for3SDC and 3SDR is shown in dashed lines in Figure 4.4(b). It can be seen
from Figure 4.4 (c) and Figure 4.4(d) that the mean values are very close; their comparison without
reflecting the effect of IM leads to erroneous conclusions. ANOVA and KW tests neglect the
variation of IM in their comparison and this ignorance leads to p–values higher than 0.05. On the
other hand, ANCOVA compares the treatment means after adjusting the variation in IM, i.e., what
would be the response means, if the two bridge sub-classes have the same IM. In other words,
ANCOVA tests the null hypothesis whether the ‘adjusted population’ means are equal. It can also
be formulated as whether the regression intercept of the two bridge sub-classes is equal
(Mangalathu et al. 2016a; 2017a). Under the assumption of the homogeneity of regression slopes,
the difference between the intercept means is equal to the difference between the adjusted means.
The regression lines of the curvature ductility for 3SDC and 3SDR are plotted in Figure 4.4(d).
ANCOVA yields the cross–section as significant (p–value of 0.010) because of the difference
74
between the intercepts. The adjusted and unadjusted means are equal only when the IM means are
Figure 4.5 – Data analysis for case 2: a) histogram of 2SSC, b) histogram of 2SSR, c) box
plot of 2SSC and 2SSR, and d) ANCOVA regression lines of 2SSC and 2SSR.
The two-span seat-type abutment bridges grouped by the cross–section shape based on the
transverse abutment response are 1) with circular cross–sections (hereafter, 2SSC) and (2) with
rectangular cross–sections (hereafter, 2SSR). ANOVA and ANCOVA identify this cross-section
shape as a significant design attribute with p–value of 0.005 and 0.048, respectively. The KS test
shows a p–value greater than 0.015 and 0.016 for 2SSC and 2SSR, and hence fails to reject the
null hypothesis that they are normally distributed. As seen from Figure 4.5(c) and Figure 4.5(d),
there is a statistically significant difference between the mean values, and both ANOVA and
ANCOVA method can capture this difference. The reason why the KW test suggests it as non-
75
significant might be due to the limited power when the data is normally distributed, which has also
been pointed out by other researchers (Vickers 2005; Kvam and Vidakovic 2007).
Figure 4.6 – Data analysis for case 3: a) histogram of 2SSS, b) histogram of 2SSP, c) box
plot of 2SSS and 2SSP, and d) ANCOVA regression lines of 2SSS and 2SSP.
The two span seat-type abutment bridges are categorized by the abutment type based on
the passive response of the abutments (1) with spread abutments (hereafter, 2SSS) and (2) with
abutment on piles (hereafter, 2SSP). The results (Figure 4.6) underscore the importance of
significant design attribute, while other grouping methods suggest it as non–significant. The mean
value of IM (in a logarithmic scale) for the passive abutment action of 2SSS and 2SSP is –1.31
and –1.13, respectively, and the comparison of the adjusted means identifies this attribute as non–
significant (Figure 4.6d). Although ANOVA identifies it as insignificant, the KW test fails here.
76
All of the above cases identify the statistical power of ANCOVA to provide an unbiased
estimator compared to ANOVA and KW methods by accounting for the variation in IMs. Although
not shown here, similar conclusions are also drawn for the other cases. To evaluate whether the
bridge sub-classes identified by ANCOVA yield a distinct response, fragility curves are developed
The following points can be inferred from the comparison of various grouping techniques:
1. ANCOVA compares the linear regression between the component response and the
models (PSDMs). The PSDMs are one of the core steps in the generation of fragility curves.
Thus, ANCOVA compares the PSDMs and the group’s bridge sub-classes based on the
difference in PSDMs. ANOVA compares the responses based on mean value of the
responses.
2. ANOVA and ANCOVA yield similar results when the mean value of ground motions
associated with the groups to be compared is the same. If there is a variation in the IMs
mean value, ANCOVA is more likely to catch the significant parameter than ANOVA.
3. The normality assumptions in ANOVA and ANCOVA seem to be satisfied in the case of
seismic responses of bridges. Hence, the statistical power in identifying the significant
attributes is greater in ANOVA and ANCOVA than the KW. It seems from the current
study that the KW is not a reliable performance based grouping approach for bridges.
5. Where the ground motions associated with the bridge groups to be compared are the same,
77
In light of the inferences from the various grouping strategies, the current study adopts
ANOVA from the point of pair-wise comparisons and the extent of numerical simulations needed
Step 2: Using Latin Hypercube Sampling method, select N ground motions from the suite of
ground motions assembled for the fragility analysis. The ground motions are selected based
Step 3: Analyze each bridge configuration in OpenSees for the selected N ground motions.
Step 4: Collect output of interest (or response) including curvature ductility, bearing displacement,
Step 5: Conduct an ANOVA to evaluate the sensitivity of each component to the variation in
bridge configurations. The results can be inferred more easily though p–value. The p–value
is the evidence against a null hypothesis or the probability that the variation between groups
occurred by chance. The p–value can be interpreted as the probability of such an 'extreme'
Step 6: Perform Fischer Method on the ANOVA output to group the bridge configurations that
have a statistically similar response. The Fisher Method compares all pairs of groups while
controlling the individual error rate. It identifies the group with the highest sensitivity and
checks a null hypothesis whether the mean values of other groups match with the most
sensitive one. If there is a match, they will be grouped together. If not, it will check the
group with the second highest sensitivity and check whether the mean value of the
78
remaining groups matches the group with second highest sensitivity. The procedure is
group the bridge classes, the uncertain parameters in Table 4.1 are kept at their mean values.
Nonlinear time history analysis (NLTHA) is carried out for the bridge models for the selected
ground motions and the maximum response of the various bridge components are recorded. Thirty
ground motions are selected from the expanded suite of 320 ground motions by Latin Hypercube
Sampling (LHS) for the grouping of bridge classes and is chosen based on a sensitivity study. The
histogram of the peak ground acceleration (PGA) and the acceleration response spectrum of the
Figure 4.7 – a) Histogram of the PGA values of the ground motion suite, b) Acceleration
response spectrum of the ground motion suite.
Various demand parameters such as column curvature ductility (μϕ, -), passive abutment
displacement (δp, mm), active abutment displacement (δa, mm), transverse abutment displacement
79
(δt , mm), bearing displacement (δb, mm), and superstructure unseating (c, mm) are used to group
the bridge classes. As mentioned before, the results are inferred in terms of p–value. A smaller p–
value refers to stronger evidence for rejecting the null hypothesis (H1). If the p–value is less than
0.05, it can be concluded that not all of the population means are equal. The sensitivity of the
seismic demand on two-span bridge configurations to the various design eras and bent
configurations are evaluated using ANOVA and is given in Table 4.3. It is clearly observed from
Table 4.3 that all of the demand parameters in rigid abutments are highly sensitive to the design
eras and bent configurations, and hence cannot be grouped together. In the case of seat abutments,
column curvature ductility and bearing displacement are the most sensitive parameters affected by
the design eras and bent configurations. The Fischer Method is carried out on ANOVA to group
the bridges that have similar seismic demands; those results are shown in Table 4.3.
Table 4.4 presents the Fischer Method grouping results for two span box-girder bridges
and Table 4.5 shows the grouping results of Era 11 bridges to number of spans. The inferences
obtained from the sensitivity study results, presented in Table 4.4 and Table 4.5, are summarized
below:
1. For the two-span seat and diaphragm abutment bridges, Era 11 shows behavior that is
distinct from the other design eras (Table 4.4). It can be inferred that the changes in the
seismic design philosophy from Era 22 to Era 33 don’t significantly change the seismic
80
2. In the case of two-span bridge configurations, the seismic demand of columns (μϕ) is the
component greatly influenced by the design eras and number of columns per bent. It
3. By comparing the seismic demand on abutments, it can be inferred that the bridge design
philosophy and number of columns per bent have more influence on the diaphragm
abutment bridge than the seat abutment bridge. It might be due to the integral connection
of the diaphragm abutment bridges at the ends which causes the abutment to share a
significant portion of the seismic demand. In the case of seat abutment bridges, the seismic
demand on the abutments is less influenced by the design eras and number of columns per
bent.
4. The seismic demand of single–column bent, two– column bent, and multi–column bent
(bent with greater than two columns) bridges are statistically different from the selected
two span bridge configurations (Table 4.5) and thus cannot be grouped together from a
5. In the case of diaphragm abutment bridges, two-span bridges have distinct seismic
demands that are distinct from three-to six-span bridge configurations for all bridge
components. Although not shown here, similar conclusions have also been noted for
6. The bridges with diaphragm abutments and seat abutments have different seismic demand
81
Table 4.4 – Results of the grouping for two-span box girder bridges.
Column ductility Abutment passive Abutment active Abutment Bearing/ Unseating
Bridge configurations transverse displacement
Mean* Grouping# Mean Grouping Mean Grouping Mean Grouping Mean Grouping
Era 11 - 1 column bent * 0.86
value B C 0.00 B 0.08 Method
B 0.75 MethodC – – –
Era 11 - 2 column bent (µ)
1.96 A 0.39 A 0.45 A 1.58 A – – –
– – –
Diaphragm abutments
*Mean is shown in logarithmic scale. #Bridge configurations that do not share common alphabet cannot be grouped together as the seismic demand of these bridges are statistically
different.
82
Table 4.5 – Results of the grouping for multi-span bridges
Column ductility Abutment passive Abutment active Abutment Bearing/Unseating
Bridge configurations transverse displacement
Mean* Grouping# Mean Grouping Mean Grouping Mean Grouping Mean Grouping
Era 11 – 2 span * 1.26
value A -0.02 A 0.01 AMethod -0.05 AMethod 0.93 A
Era 11 – 3 span (µ)
0.82 A B -0.49 A B -0.47 A B -0.70 A B 0.61 A B
Era 11 – 4 span 0.55 B -0.88 B -0.87 B -0.85 B 0.40 B b B
Era 11 – 5 span 0.53 B -0.99 B -0.97 B -0.97 B 0.36 B
Era 11 – 6 span 0.52 B -1.04 B C -1.02 B -1.03 B 0.36 C
B
*Mean is shown in logarithmic scale. #Bridge configurations that do not share common alphabet cannot be grouped together as the seismic demand of these bridges are statistically
different.
83
It is noteworthy to mention that only the abutment types, design code eras, interior
support and number of columns per bent, column cross-section, span-range and frame-
system are adopted as design attributes in this study to group bridge classes. These
attributes are selected based on the current sensitivity study, insights from the previous
research on the bridge’s seismic responses for regional risk assessment (Shinozuka et al.
2000; Mackie and Stojadinovic 2001; Avsar et al. 2001; Choi 2002; Nielson 2005; Padgett
2007; Ramanathan 2012; Moschonas et al. 2008; Banerjee and Shinozuka 2008; Mehr and
Zaghi 2016; Managalathu et al 2016a, 2016b; Zelaschi et al. 2016) and the input from
Caltrans (Caltrans, 2017). Figure 4.8 shows the proposed grouping; classification is carried
out based on the abutment type, column cross–section, pier type, number of spans, span
• Abutment types: It has been noted that the response of rigid abutments is different
• Column cross-section: On the basis of the column cross-section shape, the bridges are
classified into circular, rectangular, and oblong bridges. Soleimani et al. (2017) showed
different demands for bridges with circular, rectangular, and oblong cross sections.
• Interior support and number of columns per bent (Pier type): Sensitivity results
showed that the number of columns greater than three in a multi-column support does
not significantly impact response and therefore does need to be considered as separate
classes. The responses for bridge models having from 3-5 columns-per-bent were most
consistent (MCB, hereafter) and can be grouped together. Responses for 2-column
bents (TCB, hereafter) were statistically different from MCB and hence grouped
was shown to be distinct for each era and therefore cannot be grouped with other
support systems. Although pier-wall supports (PW, hereafter) were not explicitly
84
modelled, engineering judgment indicated this support system would also yield distinct
responses.
• Design code era: Sensitivity results showed that bridges built or rebuilt within either
of the two later design code eras (i.e. Era 22 and Era 33) had statistically similar
responses and could be grouped as ‘E22/E33’ Era for purposes of establishing demand
models (note: capacity models, particularly for the columns, are different for these eras
and hence the fragilities). Era 11 bridges were shown to have distinct response (also
capacity) and require the development of separate demand models. Mixed-era bridges
• Span range: Single-span bridges need to be treated as a separate class due to their
single-frame systems having more spans (from 2-span to 6-span) to determine if any
range could be grouped. Depending on other factors such as era and abutment, the
responses varied from being similar for all span ranges to being distinct for various
treated separately, while span-groups of two (i.e. three- and four-span bridges (S34,
hereafter), and spans greater than 4 (S4x, hereafter) were adopted for longer bridges.
framing configurations. Responses for 2-frame systems were clearly unique, but the
distinctions between bridges with a higher numbers of frames were less clear. The ‘two-
frame’ system was therefore retained as a distinct class. Larger numbers of frames such
as three frames four frames et al., are combined into the ‘multi frame’ It is noteworthy
to mention that very less number of multiple frame system are noted from the extensive
85
4.8 Conclusion
structures, as it is cumbersome and time consuming to generate fragility curves for each
structure in a specific region. Also the generation of each structure-specific fragility curve
is not warranted as some structures have similar performance or fragilities. Currently, the
addressed by performance based grouping techniques, which lead to more reliable sub-
classes of bridges relative to the traditional subjective lumping of bridges. The current
study explores the various performance based grouping techniques such as analysis of
Wallis test (KW) for the grouping of structures of similar performance. The selected
grouping methods have different underlying assumptions and approaches in grouping the
structures. The comparison of various grouping techniques is carried out with the case
study of two span and three span concrete box girder bridges in California with seat-type
analysis of variance (ANOVA) is suggested in this chapter. Thirty ground motions are
selected by Latin hypercube sampling from the suite of ground motions assembled for
California for the nonlinear time history analysis (NLTHA) of bridge models. Consistent
with the ground motions, 30 three-dimensional bridge models are created in OpenSees and
the maximum component responses are noted for each NLTHA. ANOVA is carried out for
the recorded maximum bridge responses to determine if the mean responses are statistically
similar. If they are not, the bridges are grouped by pairwise comparison using the Fischer
method.
86
The insights from the performance based grouping method, Caltrans engineering
population size of the sets of bridges sharing a particular combinations of design features
87
CHAPTER 5 CALIFORNIA BRIDGE INVENTORY
inventory. This chapter presents an in-depth study of the California bridge inventory using
(Caltrans) engineers.
historic 1971 San Fernando and the 1989 Loma Prieta earthquakes. Based on the unique
design attributes and the evolution in seismic design philosophy, California bridges can be
separated by three design eras: Era 11 (pre-1971), Era 22 (1971-1990), and Era 33 (post-
1990) (Ramanathan, 2012). In Era 11 seismic design philosophy, seismic forces were
proportional to the dead weight of the structure. Bridges were designed for a lateral seismic
force equal to 6% of the dead weight of the structures. Column shear failure and pull-out
of the longitudinal reinforcement were predominant due to the lack of ductility, as was
revealed in the 1971 San Fernando earthquake. After the 1971 San Fernando earthquake,
capacity design principles were added to the design standards. The lateral load-carrying
capacity of the bridges was increased by a factor of 2 or 2.5 and the aspects of fault
proximity, site conditions, dynamic structural response, and ductile details were considered
in the design of bridge columns. However, column shear failure in the plastic hinge regions
was typical in Era 22 due to the lack of confinement in this zone. The extensive damage
from the 1989 Loma Prieta earthquake forced Caltrans to solicit the Applied Technology
Council (ATC) to conduct a detailed study and to provide recommendations for design
standards, performance criteria, and practices. The recommendations from ATC described
in ATC-32 were incorporated in the Caltrans design manuals and led to the Era 33 design
88
or capacity design approach, which ensures a ductile failure mode in the columns. A
detailed review of the design details pertinent to various design eras is presented in this
chapter.
BIRIS is the bridge inventory assembled by Caltrans engineers for the purpose of
having a unified database for bridges; it contains bridge types, materials, operational
conditions, geometric data, functional details and other data. Table 5.1 shows the
distribution of various bridge classes in California obtained from BIRIS. Bridges are
classified based on the material and the bridge type. Box girder bridges account for the
majority of the California bridge inventory. The current study is limited to concrete box
girder, which accounts for more 30% of the California bridge inventory.
Table 5.1 – Bridge classes in California and their proportion in the overall
inventory.
Materials Total
Concrete Steel Mixed
Box girder 7839 23 166 8028
Tee Girder 2901 0 16 2917
I girder 1015 2133 603 3751
Slab 5703 15 64 5782
Culvert 3307 264 22 3593
Others 837 378 798 2013
Total: 26084
The various components of a three-span box girder bridge are illustrated in Figure
5.1. An extensive plan review was carried out to identify the design attributes and are
explained below:
89
Figure 5.1 – Illustration of major bridge components.
5.3 Abutments
Abutments can be classified into two basic types: seat abutments and diaphragm
abutments (Ramanathan 2012). Diaphragm abutments are cast monolithic with the
superstructure. As the diaphragm abutments readily engage the backfill soil during the
seismic action, it provides a great source of energy dissipation and reduces the likelihood
which is restrained longitudinally by the abutment backwall and transversely by the shear
key. Mixed abutments are supported by a diaphragm abutment at one end and seat abutment
at other end. The distribution of the abutments for various design eras are shown in Figure
5.2. Rigid abutments are the most common type in Era 11, while seat abutments are more
common in Era 33. Era 22 contains an approximately equal mix of rigid and seat abutments.
90
13% 1%
9%
45%
86% 46%
Era 22
Era 11
1% 7%
Diaphragm
Seat
92% Mixed
Era 33
Based on the support type, the abutment can be (1) on piles, (2) on spread footing
or (3) cantilever type. The distribution of the abutment based on the support type is shown
in Figure 5.3. Abutments resting on piles are the major configuration for both diaphragm
91
Diaphragm abutments (Era 11) Seat abutments (Era 11)
5%
Resting Resting
on piles on piles
31%
34% Spread 44% Spread
61% footing footing
Cantileve Cantilev
r type 25%
er type
0% 8%
Resting on Resting
Piles on Piles
Spread 46% Spread
100% footing Footing
46%
Cantilever Cantilev
type er type
92
Table 5.2 shows the statistics summary of abutment parameters for various design
eras and abutment types. The development of abutment parameter values was based on
manual review of bridge details found in plans downloaded through the BIRIS search. It is
seen from Table 5.2 that the abutment backwall height varies depending on the design era,
abutment type, and the abutment support type. Caltrans 2014 draft of bridge design aids
(BDA, hereafter) on ‘Permissible Horizontal Loads for Standard Plan and Steel HP Piles’
(Caltrans, 2017) was used to establish a typical value and representative range of pile
capacity for various types of standard piles in both sands and clays. BDA defines the
approximately 25 kips per pile was determined from BDA for a 5-foot cutoff, and the
representative range of 15-50 kips per pile showed the variability to be approximately a
factor of two above and below the central value. These ranges of values are used to derive
the abutment pile stiffness and are given in Table 5.2. To derive the coefficient of friction
were considered (Potyondy, 1961), and the range is modeled as a normal distribution with
mean of 0.40 and standard deviation of 0.075. A slip as little as 0.04-0.10 inch can mobilize
the full friction in an abutment on spread footing and hence yield displacement is assumed
to be uniformly distributed with an upper bound of 0.10 in. and lower bound of 0.04 in.
93
Table 5.2 – Distribution of parameters for abutments.
Parameter Design Type of abutment Units Distribution
era Type Parameters† Lower Upper
§ Mea Standard bound Bound
n () Deviatio (L) (U)
n ()
On piles feet LN 2.35 0.15 8.0 14.0
Diaphragm On spread feet LN 2.20 0.35 4.5 18.0
Cantilever feet U 25.0 8.33 20.0 30.0
On piles feet LN 1.95 0.20 5.0 10.5
Era11
Seat On spread feet LN 1.95 0.20 5.0 10.5
Cantilever feet U 25.0 8.33 20.0 30.0
On piles feet LN 2.39 0.20 6.5 13.0
Diaphragm On spread feet LN 2.37 0.09 9.5 12.5
Abutment
Cantilever feet - - - - -
backwall
On piles feet LN 2.45 0.18 9.5 20.0
height Era22
Seat On spread feet LN 2.50 0.09 10.5 14.5
Cantilever feet - - - - -
On piles feet LN 2.45 0.18 9.5 20.0
Diaphragm On spread feet - - - - -
Cantilever feet - - - - -
On piles feet LN 2.63 0.22 10.5 23.5
Era33
Seat On spread feet LN 2.58 0.14 11.0 19.0
Cantilever feet - - - - -
Abutment pile
All
stiffness Diaphragm On piles kip/ft LN 1.79 0.35 2.5 12.0
Eras
(lateral
capacity All
Seat On piles kip/ft LN 2.08 0.35 4.0 16.0
/deck width) Eras
Coefficient of All Diaphragm
On spread _ N 0.40 0.075 0.25 0.55
friction Eras /Seat
Yield All Diaphragm
On spread in. U 0.75 0.02 0.50 1.0
displacement Eras /Seat
Abutment
All Diaphragm
backfill soil All types B Equally split among all simulations
Eras /Seat
(clay vs sand)
§
N = normal, LN = lognormal, U = uniform, and B = Bernoulli distribution.
† and are the parameters of the distribution. These denote mean and standard deviation for a normal and uniform distribution, and mean
and standard deviation of the associated normal distribution (in log space) in the case of a lognormal distribution.
94
Table 5.3 – Distribution of parameters for bearings.
Design Type of Parameter Units Distribution
era bearings Type§ Parameters† Lower Upper
Mean Standa bound Bound
() rd (L) (U)
Deviati
on ()
Coefficient of friction
- N 0.04 0.01 0.02 0.06
(longitudinal direction)
Rocker
Coefficient of friction
- N 0.10 0.02 0.06 0.14
(transverse direction)
Stiffness per feet of
kip/in/ft LN 0.40 0.35 0.70 3.0
Era11 Elastomeric deck width
Coefficient of friction - N 0.30 0.10 0.10 0.50
Coefficient of friction - U 0.50 0.03 0.20 0.80
Friction
Yield displacement In. U 0.07 0.0003 0.04 0.10
Stiffness per feet of
kip/in/ft LN 0.77 0.52 0.7 6.0
Era22 Elastomeric deck width
Coefficient of friction - N 0.30 0.10 0.10 0.50
Stiffness per feet of
kip/in/ft LN 0.00 0.45 0.4 2.5
Era33 Elastomeric deck width
Coefficient of friction - N 0.30 0.10 0.10 0.50
§
N = normal, LN = lognormal, U = uniform, and B = Bernoulli distribution.
† and are the parameters of the distribution. These denote mean and standard deviation for a normal and uniform distribution, and mean
and standard deviation of the associated normal distribution (in log space) in the case of a lognormal distribution.
5.4 Bearings
70%
Bridges with seat abutments are rest on bearings at the abutments. Era 11 consists
of various bearing such as rocker bearings, elastomeric bearings, and friction bearings
95
(Figure 3.14), while Era 22 and Era 33 consists only of elastomeric bearings. The motion
associated with elastomeric and friction bearings are based on sliding. On the other hand,
bearings of various eras is shown by percentage in Figure 5.4. The statistical distribution
The important uncertain parameters in the modeling of superstructure are the span
length, deck-width, and type of girder (reinforced or pre-stressed). The distribution of these
parameters is derived based on an extensive plan review of bridges pertinent to the design
eras. As explained in Chapter 4, bridges are grouped based on the span length as S11 (single
span), S22 (two-span), S34 (three and four spans), and S5x (spans greater than 4). Also,
Era 22 and Era 33 bridges are grouped together from a demand perspective.
Span length is a critical parameter that governs the seismic responses of bridges;
the parameters of the span length distribution for box girder bridges are given in Table 5.4.
The span length is defined as a function of the design era, type of superstructure, and the
number of spans. Interested readers are directed to the memo on the span length models for
box girder bridges (Roblee, 2016a) and a brief summary is given in this section. Some
• In general, the span value associated with reinforced concrete (RC) girder is less
• The mean span value for multi-span PC bridge models is 135-feet for 2-span
bridges and 155-feet for 3- to 6-span bridges, independent of the design era.
96
However, the mean span length of RC bridge models is not independent of design
era.
• Two-span RC bridges are somewhat longer than other span ranges for each era.
• In the case of single-span bridges, both RC and PC have somewhat shorter span
The ratio of the approach span to the main span for multi-span bridges is defined
as the span ratio and its properties are also given in Table 5.4. The span ratio model for PC
bridges has a higher mean (0.75) than that for RC bridges (0.60), but the overall range is
Deck width parameters are determined based on the extensive plan review of
bridges (Roblee 2016b) and are given in Table 5.5. Deck width distribution is a function
of the design era and number of columns per bent. Era 22 and Era 33 are combined for
deck width distribution as the plan review suggested similar trends in Era 22 and Era 33.
97
Table 5.4 – Distribution of span length and span ratio (approach span/main span) for box girder bridges.
Span length distribution Span ratio distribution
Type of
Parameters† Type Parameters†
superstructure: §
No. Mi Stand
Design Span Reinforced Unit Lower Upper Standa Lower Upper
of x Typ ard
era range (RC) or s Mean bound Bound Mean rd bound Bound
spans % e§ Devia
Pre-stressed () (L) (U) () Deviati (L) (U)
tion
(PC) on ()
()
50 RC feet N 80 25 35 130 - - - - -
S11 1
50 PC feet N 110 35 40 180 - - - - -
75 RC feet N 95 20 55 140 - - - - -
S22 2
25 PC feet N 135 35 75 230 - - - - -
3 55 RC feet N 90 25 50 160 N 0.60 0.20 0.35 1.00
Era11 S34
4 45 RC feet N 90 25 50 160 N 0.60 0.20 0.35 1.00
5 80 RC feet N 90 20 60 125 N 0.60 0.20 0.35 1.00
S5x
6 20 RC feet N 90 20 60 125 N 0.60 0.20 0.35 1.00
10 RC feet N 80 25 35 130 - - - - -
Era22
35 PC feet N 130 35 50 220 - - - - -
S11 1
15 RC feet N 105 40 35 200 - - - - -
Era 33
40 PC feet N 130 35 50 220 - - - - -
5 RC feet N 95 20 55 140 - - - - -
Era22 S22
35 PC feet N 135 35 75 230 - - - - -
2
10 RC feet N 135 35 85 200 - - - - -
Era 33 S22
50 PC feet N 135 35 75 230 - - - - -
10 RC feet N 90 25 50 160 N 0.60 0.20 0.35 1.00
3
20 PC feet N 155 45 75 250 N 0.75 0.20 0.40 1.00
Era22 S34
5 RC feet N 90 25 50 160 N 0.60 0.20 0.35 1.00
4
10 PC feet N 155 45 75 250 N 0.75 0.20 0.40 1.00
5 RC feet N 110 35 55 190 N 0.60 0.20 0.35 1.00
3
30 PC feet N 155 45 75 250 N 0.75 0.20 0.40 1.00
Era 33 S34
5 RC feet N 110 35 55 190 N 0.60 0.20 0.35 1.00
4
15 PC feet N 155 45 75 250 N 0.75 0.20 0.40 1.00
10 RC feet N 90 20 60 125 N 0.60 0.20 0.35 1.00
5
10 PC feet N 155 35 95 240 N 0.75 0.20 0.40 1.00
Era22 S5x
5 RC feet N 90 20 60 125 N 0.60 0.20 0.35 1.00
6
5 PC feet N 155 35 95 240 N 0.75 0.20 0.40 1.00
15 RC feet N 125 35 75 165 N 0.60 0.20 0.35 1.00
5
30 PC feet N 155 35 95 240 N 0.75 0.20 0.40 1.00
Era 33 S5x
5 RC feet N 125 35 75 165 N 0.60 0.20 0.35 1.00
6
20 PC feet N 155 35 95 240 N 0.75 0.20 0.40 1.00
§
N = normal, LN = lognormal, U = uniform, and B = Bernoulli distribution.
† and are the parameters of the distribution. These denote mean and standard deviation for a normal and uniform distribution, and mean and standard deviation of the associated normal
distribution (in log space) in the case of a lognormal distribution.
98
Table 5.5 – Distribution of deck width for box girder bridges.
Column Distribution Cell distribution
type (no. Parameters
Design of Lower Upper
Mix % Standard % % %
era columns) Unit Type§ Mean bound Bound Type Type Type
Deviation distr. distr. distr.
() (L) (U)
()
25 feet N 26.5 1.5 22 30 3 cell 100 - - - -
1 50 feet N 34 1.2 30 38 3 cell 70 5 cell 30 - -
25 feet N 40 1.5 38 46 3 cell 40 5 cell 60 - -
15 feet N 34 2.0 30 38 3 cell 50 5 cell 50 - -
2 25 feet N 41 5 38 48 3 cell 25 5 cell 75
Era 11 15 feet N 58 26 48 74 5 cell 25 7 cell 50 9 cell 25
10 feet N 48 18 38 56 5 cell 65 7 cell 35 - -
3 15 feet N 66 9 56 74 7 cell 50 9 cell 50 - -
5 feet N 80 9 74 92 9 cell 70 11 cell 30 - -
5 feet N 60 34 38 72 5 cell 25 7 cell 35 9 cell 40
4
10 feet N 88 34 72 106 9 cell 25 11 cell 75 - -
15 feet N 28 1.2 22 30 3 cell 100 - - - -
20 feet N 34 4 30 38 3 cell 85 5 cell 15
1
55 feet N 42 2 38 46 3 cell 75 5 cell 25 - -
10 feet N 50 14 46 60 3 cell 30 5 cell 50 7 cell 20
20 feet N 43 7 36 50 3 cell 40 5 cell 60
Era
2 15 feet N 57 8 50 66 5 cell 80 7 cell 20
22/Era
10 feet N 73 22 66 88 5 cell 25 7 cell 50 9 cell 25
33
10 feet N 59 18 50 68 5 cell 50 7 cell 50
3 15 feet N 79 20 68 88 7 cell 50 9 cell 50
10 feet N 98 20 88 108 7 cell 20 9 cell 40 11 cell 40
5 feet N 75 32 58 90 5 cell 25 7 cell 40 9 cell 35
4
15 feet feet 107 38 90 128 9 cell 40 11 cell 35 13 cell 25
§
N = normal, LN = lognormal, U = uniform, and B = Bernoulli distribution.
† and are the parameters of the distribution. These denote mean and standard deviation for a normal and uniform distribution, and mean and standard deviation of the associated normal
distribution (in log space) in the case of a lognormal distribution
99
5.5.3 Deck cross-section properties
Figure 5.5 illustrates the typical cross-section of box girder bridges. The height of
the box girder is a function of the span length; the acceptable depth-to span ratios are 0.055
and 0.04 for RC and PC concrete boxes, respectively. The cross-section details noted from
the plan review are presented in Table 5.4. It has been noted that the wall thickness (twall)
is constant across the design eras and type of superstructure (RC or PC). A constant value
is adopted for the bottom flange thickness (tbot) for each design era and is given in Table
5.6. The top flange thickness is a function of the center-to-center spacing of the girders
100
Table 5.7 – Box-girder deck slab thickness (MTD, 2008).
101
5.6 Columns
As mentioned before, the column details vary depending on the design era due to
the changes in design philosophy. The design philosophies adopted for California bridges
have been significantly influenced by the historic 1971 San Fernando and the 1989 Loma
Prieta earthquakes. In seismic design philosophy during Era 11, seismic forces were
proportional to the dead weight of structures. Bridges were designed to resist a lateral
seismic force equal to 6% of the dead weight of the structures. After the 1971 San Fernando
earthquake, capacity design principles were introduced in the seismic design standards.
The lateral load-carrying capacity of the seismically designed bridges increased by a factor
of 2 or 2.5. In addition, several aspects, including fault proximity, site conditions, dynamic
structural response, and ductile details were considered in the design of bridge columns.
The extensive damage from the 1989 Loma Prieta earthquake forced Caltrans to solicit the
leading to the Era 33 columns. The design attributes and statistical properties of bridge
columns are identified by an in-depth review of bridge plans pertinent to the design eras
In this study, column heights are measured as the height between the underside of
the bridge deck and the top of column footing by manual plan review. The basic statistics
of the column height are provided in Table 5.8. As noted in Table 5.8, the median column
102
Table 5.8 – Column height distribution.
Design Units Distribution
era Type Parameters† Median Lower Upper
§
Mean () Standard Deviation () bound (L) Bound (U)
Era11 ft LN 3.06 0.13 21.5 16.5 28.0
Era 22 ft LN 3.14 0.16 23.2 17.0 32.0
Era 33 ft LN 3.22 0.18 25.0 17.5 36.0
§
N = normal, LN = lognormal, U = uniform, and B = Bernoulli distribution.
† and are the parameters of the distribution. These denote mean and standard deviation for a normal and uniform distribution, and
mean and standard deviation of the associated normal distribution (in log space) in the case of a lognormal distribution
Various cross sections such as circular, rectangular, and oblong are noted from the
plan review. Era 11 consists of circular and rectangular cross-sections while Era 22 and
Era 33 contain circular and oblong cross-sections. As seen in Figure 5.6, where the cross-
section details of various design eras are presented, each era consists of a wide range of
cross-sections. The percentage distribution of the Era 11 columns is given in Table 5.9.
103
Table 5.9 – Distribution of column cross-sections.
Column
Design Type of Diameter Length Breadth
cross- Mix %
era bent (D, in.)* (Lc, in.)* (Bc, in.)*
section
48 - - 10
Circular 60 - - 20
72 - - 20
Single - 72 36 10
- 96 36 15
Rectangular
- 120 36 10
- 96 48 15
Era 11
36 - - 15
Circular 48 - - 20
60 - - 5
Multi - 42 30 20
- 48 36 25
Rectangular
- 60 42 10
- 96 36 5
60 - - 10
66 - - 25
Circular
72 - - 15
Single 84 - - 10
- 72 48 10
Oblong - 96 48 15
Era 22 - 99 66 15
48 - - 35
60 - - 10
Circular
66 - - 20
Multi
72 - - 10
- 60 36 10
Oblong - 72 48 15
60 - - 5
66 - - 15
Circular
84 - - 20
Single 108 - - 10
- 72 48 5
- 96 48 10
Oblong
- 99 66 25
- 108 72 10
Era 33 48 - - 30
60 - - 10
Circular
66 - - 25
Multi 84 - - 5
- 72 48 15
Oblong - 99 66 10
- 126 84 5
*Refer figure 5.6
104
Figure 5.6 – Column cross-sections for various design eras of box girder bridges.
105
5.6.3 Column material properties
The bridge classes considered in this study use concrete as the construction material
and the statistical properties of the concrete are given in Table 5.10. Following the
normal distribution. The statistical properties of the yield strength of the reinforcing steel
The statistical properties of the column reinforcement identified from the plan
review of bridges are given in Table 5.11. In Era 11, the column shear reinforcement
consisted of #4 transverse stirrups spaced at 12 in. on center regardless of the column size
or the size of the longitudinal reinforcing bars. Uniform distribution is assumed for the
106
Table 5.11 – Statistical distribution of column reinforcement details.
Distribution
Parameters†
Design Lower Upper
Parameter Units Standard
era Type§ Mean bound Bound
Deviation
() (L) (U)
()
Longitudinal steel
All - U 2.00 0.33 1.0 3.0
reinforcement ratio
Era11 - #4 @ 12 in. irrespective of the cross-section
Transverse steel
Era 22/
reinforcement ratio - U 0.85 0.07 0.4 1.3
Era33
§
N = normal, LN = lognormal, U = uniform, and B = Bernoulli distribution.
† and are the parameters of the distribution. These denote mean and standard deviation for a normal and uniform
distribution, and mean and standard deviation of the associated normal distribution (in log space) in the case of a
lognormal distribution.
5.7 Foundations
The foundation provides a means to transmit service and ultimate loads from the
structure to the underlying soil. Foundations can be classified into two types: shallow
107
foundations and deep foundations. As the name implies, the loads from the structure are
transferred to the underlying soil at a shallow depth for shallow foundations. Deep
foundations are provided when soil conditions are not favorable to shallow foundations
and transfer the load through piles. The type of foundation for a particular bridge is
determined by various factors such as soil conditions, overhead clearance, existing utilities,
and proximity to existing facilities such as buildings and railroads. The possible types of
footings are shown in Figure 5.7 and the statistical distribution of foundations across the
design eras is shown in Figure 5.8. The rotational and translational stiffness of the
Era 33
17% Spread
footing
25%
Footing
58% on piles
Shaft
piles
108
Table 5.12 – Distribution of foundation rotational stiffness (106 kip-in/rad).
Type of Type of Foundation Transverse to bridge direction Trans/Long stiffness ratio
bent footing fixity Parameters Parameters
Design Lower Upper Lower Upper
Mix % Standard Standard
era Type§ Mean bound Bound Type§ Mean bound Bound
Deviation Deviation
() (L) (U) () (L) (U)
() ()
pile fixed LN 25.0 2.5 10 62.5 LN 1.5 1.5 1.0 2.25
single 100
spread fixed LN 25.0 2.5 10 62.5 LN 1.5 1.5 1.0 2.25
pinned 37.5 LN 2.5 2.5 1.0 6.3 LN 1.0 1.5 0.67 1.50
Era 11 pile
fixed 12.5 LN 4.0 2.5 1.6 10.0 LN 1.0 1.5 0.67 1.50
multiple
pinned 37.5 LN 2.5 2.5 1.0 6.3 LN 1.0 1.5 0.67 1.50
spread
fixed 12.5 LN 4.0 2.5 1.6 10.0 LN 1.0 1.5 0.67 1.50
pile fixed 50 LN 80.0 2.5 32.0 200.0 LN 1.5 1.5 1.0 2.25
single
spread fixed 50 LN 50.0 2.5 20.0 125.0 LN 1.3 1.3 1.0 1.70
pinned 25 LN 12.0 2.5 4.8 30.0 LN 1.0 1.5 0.67 1.5
Era 22 pile
fixed 25 LN 18.0 2.5 7.2 15.0 LN 1.0 1.5 0.67 1.5
multiple
pinned 25 LN 12.0 2.5 4.8 30.0 LN 1.0 1.5 0.67 1.5
spread
fixed 25 LN 18.0 2.5 7.2 15.0 LN 1.0 1.5 0.67 1.5
pile fixed 90 LN 190.0 2.5 76.0 475.0 LN 1.15 1.15 1.00 1.32
single
spread fixed 10 LN 50.0 2.5 20.0 125.0 LN 1.15 1.15 1.00 1.32
pinned 50 LN 20.0 2.5 8.0 50.0 LN 1.20 1.25 0.96 1.50
Era 33 pile
fixed 0 LN 30.0 2.5 12.0 75.0 LN 1.20 1.25 0.96 1.50
multiple
pinned 50 LN 20.0 2.5 8.0 50.0 LN 1.20 1.25 0.96 1.50
spread
fixed 0 LN 30.0 2.5 12.0 75.0 LN 1.20 1.25 0.96 1.50
109
Table 5.13 – Distribution of foundation translational stiffness (kip/in).
Type of Type of Foundation Transverse to bridge direction Trans/Long stiffness ratio
bent footing fixity Parameters Parameters
Design Lower Upper Lower Upper
Mix % Standard Standard
era Type§ Mean bound Bound Type§ Mean bound Bound
Deviation Deviation
() (L) (U) () (L) (U)
() ()
pile fixed LN 1250.0 2.5 500.0 3125.0 LN 1.0 1.0 1.0 1.0
single 100
spread fixed LN 1250.0 2.5 500.0 3125.0 LN 1.0 1.0 1.0 1.0
pinned 37.5 LN 625.0 2.5 250.0 1562.5 LN 1.0 1.0 1.0 1.0
Era 11 pile
fixed 12.5 LN 625.0 2.5 250.0 1562.5 LN 1.0 1.0 1.0 1.0
multiple
pinned 37.5 LN 625.0 2.5 250.0 1562.5 LN 1.0 1.0 1.0 1.0
spread
fixed 12.5 LN 625.0 2.5 250.0 1562.5 LN 1.0 1.0 1.0 1.0
pile fixed 50 LN 2000.0 2.5 800.0 5000.0 LN 1.0 1.0 1.0 1.0
single
spread fixed 40 LN 2000.0 2.5 800.0 5000.0 LN 1.0 1.0 1.0 1.0
pinned 25 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
Era 22 pile
fixed 12.5 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
multiple
pinned 25 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
spread
fixed 12.5 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
pile fixed 65 LN 2500.0 2.5 1000.0 6250.0 LN 1.0 1.0 1.0 1.0
single
spread fixed 10 LN 2500.0 2.5 1000.0 6250.0 LN 1.0 1.0 1.0 1.0
pinned 35 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
Era 33 pile
fixed 0 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
multiple
pinned 35 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
spread
fixed 0 LN 1000.0 2.5 400.0 2500.0 LN 1.0 1.0 1.0 1.0
110
5.8 Other uncertain parameters
5.8.1 Damping
Bavirisetty et al. (2003) estimated the 2nd and 98th percentile of damping ratios in bridges
to be 0.02 and 0.07 respectively. The recommendations of Feng et al. (1999) for tall buildings are
extended to bridges by the researchers (Nielson 2005; Padgett 2007) and the damping uncertainty
is modeled using a normal distribution. Based on these studies, damping is modeled as normal
distribution (mean = 4.5%, standard deviation = 1.25%) and is shown in Table 5.14.
Mass sources such as parapets and barrier rails, variable deck slab thickness, electric poles
and other equipment, re-pavement procedures, and variation in material densities are not
considered in the OpenSees bridge model; because of this, it has been decided to account for these
mass sources explicitly with the addition of a mass factor. The mass factor is assumed to be
Per Caltrans, designs of shear keys are categorized as either isolated, as new emerging
designs, or non-isolated, as older conventional designs. Since the isolated design is a new type of
design that does not appear in the existing inventory, the current study will focus only on the non-
isolated type. The non-isolated shear keys were designed to withstand dead-load reaction for 0.3
to 0.5g. However, they have been shown in reality to be on the order of 3 times stronger (Caltrans,
2017). Based on the recommendation from Caltrans, shear key is modeled using lognormal
111
5.8.4 Gap
Lognormal distribution is adopted for the gap between the deck and the superstructure with a
median value of -0.20 and standard deviation of 0.50. Further, the gap between the superstructure
and shear keys is assumed to be uniformly distributed between 0 and 1.5 in.
The ground motions considered in the present study have fault-normal and fault-parallel
components. Mackie et al. (2011) concluded that there is a negligible effect of the angle of
incidence on the mean ensemble response of bridge components and hence the incident angle is
not considered as a major source of uncertainty in the study. As such, the two horizontal
components of ground motions are assigned simultaneously to the longitudinal and transverse
direction of the bridge and the orientation is assigned randomly. The effects of vertical acceleration
5.9 Closure
This chapter presents the extensive plan review and analysis of the California box girder bridge
inventory using the BIRIS assembled by Caltrans engineers. The bridges are divided into three
112
eras: Era 11 (pre 1971), Era 22 (1971 – 1990), and Era 33 (post 1990), based on the seismic design
superstructure, and columns are gathered across the three design eras to aid in the development of
stochastic finite element models for the generation of probabilistic seismic demand models
(PSDMs) and fragility curves. Also, the bridge design details and physical characteristics
identified in this chapter help to capture the vulnerabilities associated with various components.
The input parameters are assumed to be uncorrelated in the current study and further studies are
113
CHAPTER 6 SYSTEM AND COMPONENT FRAGILITY CURVES
increased awareness of potential seismic hazards and the associated vulnerability of structures.
The evaluation of economic losses after earthquakes is primarily based on the prediction of
curves, a statistical function that gives the conditional probability of exceeding a certain damage
state given a certain ground motion intensity measure (IM). Component and system fragility curves
can be useful in prioritizing both post-earthquake emergency responses and field inspections.
Generating seismic fragility curves involves the convolution of demand models and
capacity models. This chapter explains the fragility framework, including the formulation of
probabilistic seismic demand models (PSDMs) and capacity models. The methodology presented
in this section is used in this study to develop system and component fragility curves for single
frame multi-span box-girder bridges in California. Comparisons are conducted among various
bridge classes to assess the relative vulnerability of each class, and are presented in detail in this
chapter.
The multiphase framework used by numerous researchers (Nielson 2005, Padgett 2007,
Ramanathan et al. 2015, Jeon et al. 2015) is adopted in the current study to shed light on the
fragilities of various bridge classes and the effects of various bridge components on bridge
fragilities. This methodology also helps to generate system as well as component fragilities. The
parameters listed in Chapter 5 are varied to capture uncertainties in bridge classes. Input variables
are sampled across the range of parameters presented in the Chapter 5 using Latin Hypercube
114
Sampling technique in order to generate statistically significant yet nominally identical three–
dimensional bridge models. The variables are randomly paired with the selected suite of ground
motions. The two orthogonal components of the ground motions are randomly assigned to the
longitudinal and transverse direction of the bridge axis. A set of 320 simulations of nonlinear time
history analyses (NLTHAs) is performed for all bridge-ground motion pairs to monitor the
maximum response of various bridge components. Figure 6.1 shows the schematic of the
procedure that was adopted to capture the demand of various bridge components due to the ground
motions.
Figure 6.1 – Schematic representation of the NLTHA procedure used to derive the
PSDMs.
115
Table 6.1 – Engineering demand parameters for bridge components monitored in NLTHA.
Component Engineering Notation Units
demand parameter
The current study considers the vulnerability of multiple components: columns, abutment
seat (seat type abutments), elastomeric bearings, joint seal, restrainer cables (retrofitted bridges),
deck displacement, foundations, abutments, and shear keys. The engineering demand parameters
(EDP) representing the above components are indicated in Table 6.1. The following section
Fragility curves require the convolution of the demand model and capacity models. As
mentioned before, the seismic demands on bridge components are obtained by the three-
dimensional NLTHAs of bridge models. The peak response of the components (di, e.g., column
curvature ductility, bearing deformations, and abutment deformations) is recorded for each
NLTHA. Based on Cornell et al. (2002), PSDMs are defined as the linear regression of pairs of D
and IM in the log-transformed space, as illustrated in Figure 6.2. These can be written as
116
where a and b are the regression coefficients. The coefficients a and b are obtained by
performing a linear regression analysis on D-IM pairs in the log-transformed space. Dispersion,
ln d ln ( S d )
2
i
d / IM i 1
(6.2)
N 2
The development of a probabilistic seismic demand model using the results of the NLTHA
forms the demand side of the fragility formulation. The next crucial step in the fragility formulation
is the formulation of capacity, or limit state, models. The capacity models are described by a two-
parameter lognormal distribution with median, Sc and dispersion, βc. Discrete damage states are
defined for each component corresponding to the significant change in its response and consequent
to its own performance and the performance of the bridge at both the global and system levels. A
general description of the component damage thresholds (CDT) and bridge system-level damage
states (BSST) is given in Table 6.2 and 6.3, respectively (Ramanathan et al. 2015).
117
The bridge components are categorized as either primary or secondary. Primary
components include the columns and abutment seat; these are the components that affect the
vertical stability and load carrying capacity of the bridge. Secondary components are those whose
failure will not force the closure of the bridge; this includes abutment deformations, shear key
displacement, and others. As the failure of a primary component affects the load carrying capacity
and stability of the bridge system, CDTs of primary components map directly into BSSTs. Only
two broad CDTs, CDT-0 and CDT-1, are defined for the secondary components and these map
directly into BSST-0 and BSST-1. CDTs and BSSTs were developed in close collaboration with
Caltrans. The number of components used to integrate the system fragility varies based on the
BSST under consideration. Such mapping ensures similar consequences in terms of repair and
Table 6.2– Component level damage state descriptions – Component Damage Thresholds
(CDT).
118
Table 6.3 – General description of BSSTs along with CDTs.
ShakeCast Inspection
Low Medium Medium-High High
Priority levels
Emergency Repair
Implications
Is shoring/bracing Very unlikely Unlikely Likely Very likely
needed?
Is roadway leveling Unlikely Likely Very Likely Very Likely -
needed? Detour
Component Damage
Range
Primary components CDT-0 to 1 CDT-1 to 2 CDT-2 to 3 Above CDT-3
Secondary components CDT-0 CDT-1 NA NA
A significant contribution of the present study is the suggestion of capacity limit states for
columns (CCLS) based on extensive experimental review. A review of the existing research
pertinent to various design eras was conducted to collect the experimental data for bridge columns,
and statistical analysis was carried out to suggest the CCLS for bridge columns. Such an exercise
helps to support seismic risk evaluation of bridges in California by developing a new generation
of more accurate and useful bridge fragility models for incorporation into the ShakeCast
As mentioned before, the intention of the current study is to suggest fragility relationships for
damage states that range from minor spalling of concrete to complete bridge collapse. The
definitions of these limit states and their operational consequences are given in Table 6.4.
119
Table 6.4 – General definition of column capacity limit states.
The design philosophies for California bridges have been significantly influenced by the
historic 1971 San Fernando and the 1989 Loma Prieta earthquakes. Based on unique design
attributes and evolutions in seismic design philosophy, California bridges are categorized into
three design eras: Era 11 (pre-1971), Era 22 (1971-1990), and Era 33 (post-1990). In Era 11
seismic design philosophy, seismic forces were proportional to the dead weight of structures.
Bridges were designed to resist a lateral seismic force equal to 6% of the dead weight of the
structures. Shear failure and pull-out of the longitudinal reinforcement in columns were
predominant due to the lack of ductility, as was revealed by the 1971 San Fernando earthquake.
After the 1971 San Fernando earthquake, capacity design principles were introduced in the
seismic design standards. The lateral load carrying capacity of seismically-designed bridges
increased by a factor of 2 or 2.5. In addition, several aspects, including fault proximity, site
conditions, dynamic structural response, and ductile details, were considered in the design of
120
bridge columns. However, shear failure in the columns was still observed in Era 22 columns due
The extensive damage from the 1989 Loma Prieta earthquake forced Caltrans to solicit the
Applied Technology Council (ATC) to conduct a detailed study and provide recommendations for
design standards, performance criteria, and practices. The recommendations described in ATC-32
were incorporated in Caltrans design manuals, leading to the distinctive Era 33 columns. The
fundamental emphasis in the Era 33 design philosophy was on the displacement-based or capacity
design approach, which ensures a ductile failure mode in the columns. The design details pertinent
to various design eras are given in Table 6.5, 6.6, and 6.7 respectively.
Table 6.8, 6.9, and 6.10, respectively, summarize the extensive literature review on
experimental investigations of bridge columns. The tables also present the geometric features of
the test column, and the values of displacement and curvature ductility limit states for various
levels of damage.
121
Table 6.5 – Design details for columns in Era 11 (pre-1971).
122
Table 6.6 – Design details of columns in Era 22 (1971-1990).
123
Table 6.7 – Design details for columns in Era 33 (post-1990).
124
Table 6.8 – Summary of limit states for Era 11 columns.
Section properties Displacement ductility Curvature ductility Reference
Colu
Lon
mn Colu
Colum g.
Cross- dia Length * mn Scal Design CCLS- CCLS- CCLS- CCLS- CCLS- CCLS- CCLS- CCL
n Steel
section mete breadth (in.) heigh e code 0 -1 -2 -3 -0 -1 -2 S--3
model ratio
r t (ft.)
(%)
(in.)
R-I Rectangle - 28.75 19.25 12.0 0.4 Caltrans 2.55 0.80 1.00 2.00 3.00 0.80 2.00 6.00 10.00
Sun et al.
R-5 Rectangle - 28.75 19.25 12.0 0.4 Caltrans 5.00 0.80 2.00 3.00 4.00 0.80 2.50 3.50 6.00 (1993)
SRPH-
Circular 24 - 7.0 0.4 Caltrans 5.4 0.80 1.00 1.50 2.00 0.80 1.00 1.54 2.56 Hose et al.
6
(1997)
col #2 Circular 24 - 6.0 0.4 ACI 1.06 0.80 1.50 2.00 4.00 0.80 2.26 3.52 8.56 Priestley et al.
(1996)
col-1 Circular 24 - 12 0.4 Caltrans 2.53 0.80 1.00 1.50 4.00 0.80 1.00 2.31 8.87
Chai et al.
col-3 Circular 24 - 12 0.4 Caltrans 2.53 1.00 1.50 3.00 5.00 1.00 2.31 6.25 11.49 (1991)
T1 Circular 10 - 3.33 0.28 Caltrans 2 0.80 2.00 3.00 5.00 0.80 2.93 4.85 8.71
Jaradat et al.
T2 Circular 10 - 3.33 0.28 Caltrans 1.1 0.80 2.00 4.00 5.00 0.80 2.93 6.78 8.71 (1998)
T3 Circular 10 - 3.33 0.28 Caltrans 2 0.80 2.00 3.00 4.00 0.80 2.93 4.85 6.78
S1 Circular 10 - 5.83 0.28 Caltrans 2 0.80 1.00 1.50 2.00 0.80 1.00 2.33 3.65
S2 Circular 10 - 5.83 0.28 Caltrans 1.1 0.80 2.00 3.00 5.00 0.80 3.65 6.31 11.62
S3 Circular 10 - 5.83 0.28 Caltrans 1.1 1.00 2.00 3.00 4.00 1.00 3.65 6.31 8.96
S1 Circular 20 - 5.0 0.33 Washington 0.99 0.8 1.82 5.04 8.27 0.8 2.48 8.28 14.10
Ranf et al.
S3 Circular 20 - 5.0 0.33 Washington 0.99 0.8 1.34 2.45 3.56 0.8 1.61 3.61 5.61 (2006)
S15 Circular 20 - 5.0 0.33 Washington 0.99 0.8 2.05 4.03 6 0.8 2.89 6.46 10.01
C2 Circular 20 - 5.0 0.33 Washington 0.99 0.8 1.47 2.65 3.83 0.8 1.85 3.97 6.10
C4 Circular 20 - 5.0 0.33 Washington 0.99 0.8 1.52 4.6 7.68 0.8 1.94 7.49 13.04
C3R Circular 20 - 5.0 0.33 Washington 0.99 0.8 1.32 3.51 5.7 0.8 1.58 5.52 9.47
125
Table 6.9 – Summary of limit states for Era 22 columns.
Section properties Displacement ductility Curvature ductility Reference
Colu Lengt
Colu Long.
mn h*
Column Cross- mn Design Steel CCLS- CCLS- CCLS- CCLS- CCLS- CCLS- CCLS- CCL
diam bread Scale
model section height code ratio -0 -1 -2 -3 -0 -1 -2 S--3
eter th
(ft.) (%)
(in.) (in.)
SRPH-1 Circular 24 - - 0.4 Caltrans 2.7 0.77 1.00 6.00 8.00 0.77 2.77 7.00 9.26 Hose et al.
SRPH-2 Circular 24 - - 0.4 Caltrans 5.4 0.72 1.50 3.00 4.50 0.72 1.60 4.15 7.80 (1997)
SRPH-3 Circular 24 - - 0.4 Caltrans 5.4 0.69 2.50 5.00 7.30 0.69 3.10 5.80 9.23
SRPH-7 Circular 24 - - 0.4 Caltrans 5.4 0.75 2.00 4.00 6.00 0.75 1.60 4.56 7.32
Sanchez
MG-2 Circular 28.8 - 8.8 0.4 Caltrans 2.0 0.86 2.3 2.88 3.45 0.86 4.48 6.04 7.57 et al.
24 (1997)
RDS-2 Oblong - 13.0 0.4 Caltrans 2.0 2.0 6.0 8.0 10.0 3.75 14.77 20.28 25.79
36
Asad and
PEER- Xiao
Circular 16 - 6.0 0.33 Caltrans 1.17 1.3 5.08 4.16 5.55 1.75 11.27 8.95 12.46
COL-1 (2005)
PEER-
Circular 16 - 6.0 0.33 Caltrans 1.17 1.11 2.80 4.17 5.55 1.277 5.534 8.984 12.46
COL-3
Stone and
Flexure Circular 60 - 30.0 1.0 Caltrans 2.0 1.0 3.00 5.00 6.60 3.87 6.74 12.49 17.08 Cheok
Shear Circular 60 - 15.0 1.0 Caltrans 2.0 1.0 4.00 6.00 10.00 3.09 7.26 11.44 19.79 (1989)
126
Table 6.10 – Summary of limit states for Era 33 columns.
Section properties Displacement ductility Curvature ductility
Colu Long.
Column Length *
Column Cross- mn Scal Design Steel CCLS- CCLS- CCL CCL CCLS- CCLS- CCL CCL Reference
diamete breadth
model section heigh e code ratio -0 -1 S--2 S--3 -0 -1 S--2 S--3
r (in.) (in.)
t (ft.) (%)
328 Circular 24 - 6.0 0.5 Caltrans 2.72 1.40 3.51 4.91 9.12 2.07 7.64 11.36 22.51
328T Circular 24 - 6.0 0.5 Caltrans 2.72 1.57 4.71 7.84 10.20 2.51 10.81 19.12 25.35 Calderone et
al. (2001)
828 Circular 24 - 16.0 0.5 Caltrans 2.72 1.26 4.41 4.41 7.77 1.91 12.90 12.90 24.63
1028 Circular 24 - 20.0 0.5 Caltrans 2.72 1.81 5.81 6.46 9.04 3.94 18.47 20.82 30.19
ISL 1.0 Oblong - 12 17.5 4.83 0.2 Caltrans 2.00 0.8 1.5 5.6 9.6 0.50 2.24 12.40 22.31
ISL 1.5 Oblong - 12 20.25 6 0.2 Caltrans 2.00 1.5 2.4 7.5 10.4 2.35 4.78 18.53 26.35
ISH 1.0 Oblong - 10 14.5 7.62 0.2 Caltrans 2.90 0.9 1.4 3.6 4.7 1.00 2.19 8.72 11.99 Correal. et al.
(2007)
ISH 1.25 Oblong - 10 16.75 8.15 0.2 Caltrans 2.80 0.7 1.4 3.7 4.7 1.00 2.22 9.22 12.26
ISH 1.5 Oblong - 10 15.62 8.79 0.2 Caltrans 2.90 1 1.6 2.2 4.7 1.00 2.86 4.72 12.46
ISH1.5T Oblong - 10 16.75 8.79 0.2 Caltrans 2.90 1.0 1.7 2.8 3.8 1.00 3.18 6.60 9.71
415 Circular 24 - 8.0 0.33 Caltrans 1.5 1.0 2.7 5.0 8.0 1.00 5.04 10.59 17.78
407 Circular 24 - 8.0 0.33 Caltrans 0.75 1.0 2.5 3.0 6.0 1.00 4.60 5.80 12.99
Lehman and
430 Circular 24 - 8.0 0.33 Caltrans 1.5 1.0 3.0 5.0 7.0 1.00 5.88 10.59 15.41 Moehle
(2000)
815 Circular 24 - 8.0 0.33 Caltrans 1.5 1.0 2.0 3.0 5.0 1.00 3.47 5.94 10.89
1015 Circular 24 - 8.0 0.33 Caltrans 1.5 1.0 2.0 3.0 5.0 1.00 3.47 5.94 10.89
Hose et al.
SRPH-4 Circular 24 - 7.0 0.4 Caltrans 5.4 1.00 2.50 8.00 - 1.00 2.00 14.60 - (1997)
Orozco et al.
VP-2 Circular 16 - 6.0 0.4 Caltrans 1.17 1.0 3.4 7.0 8.3 1.00 5.10 10.80 13.85
(1999)
RDS-1 Oblong - 24 ´ 36 13 0.4 Caltrans 1.64 2.0 4.0 8.0 12.0 2.30 4.50 10.90 17.30 Sanchez et al.
(1997)
RDS-6 Oblong - 24 ´ 36 13 0.4 Caltrans 2.00 2.0 6.0 8.0 12.0 1.50 5.30 11.40 17.40
H/D (6) Circular 24 - 12.0 0.5 Caltrans 2.10 0.75 4.50 6.0 18.0 0.75 8.71 12.01 38.43
Shanmugam
(2009)
H/D (3) Circular 24 - 6.0 0.5 Caltrans 2.10 0.75 4.33 13.75 17.0 0.75 5.03 16.42 20.36
127
Table 6.11 presents the summary of the ductility values for various design eras.
Column curvature ductility (µ) is chosen as the EDP for columns and the median values
Table 6.11– Statistical summary of ductility values for various design eras.
6.1.3 Abutments
study to account for the unseating potential. Bridge seat widths chronologically increased
from the 4 – 12 inch range in Era 11, to the 12 – 24 inch range in Era 22, to greater than 24
inches in Era 33. Table 6.12 gives the median CDT values for abutment seats for various
128
design eras. The values were developed based on previous studies (Ramanathan, 2012) and
The CDT values for other components such as superstructure deck, abutment
shear keys, and joint seals are consistent with previous research conducted by Ramanathan
(2012). Interested readers are directed to Ramanathan (2012) for a more detailed
explanation of the CDT values of these components. Table 6.13 provides the summary
values of the CDT and CCLS value for various bridge components. The capacity models
βc. βc is assigned as 0.35 in a subjective manner due to lack of sufficient information and
is adopted as the same across the components and the respective damage states.
129
Table 6.13– Summary of CDT values for various bridge components.
Components EDP Units Median value, Sc c
CDT- CDT- CDT- CDT
0 1 2 -2
Primary Components
Columns
Era 11 Curvature NA 0.8 2.0 5.0 8.0 0.35
ductility
Era 22 Curvature NA 1.0 5.0 8.0 11.0 0.35
ductility
Era 33 Curvature NA 1.0 5.0 11.0 17.0 0.35
ductility
Abutment seat
Era 11 Displacement Inches 0.5 1.0 2.0 3.0 0.35
Era 22 Displacement Inches 1.0 4.5 10.0 15.0 0.35
Era 33 Displacement Inches 1.5 4.5 14.0 21.0 0.35
Secondary Components
Joint Seal Displacement Inches 2.0 5.0 NA NA 0.35
Bearings Displacement Inches 1.0 4.0 NA NA 0.35
Shear keys Displacement Inches 1.0 5.0 NA NA 0.35
Deck Displacement Inches 4.0 12.0 NA NA 0.35
Bent foundation
Translation Displacement Inches 1.0 4.0 NA NA 0.35
Rotation Rotation Radian 1.5 6.0 NA NA 0.35
Abutments
Passive Displacement Inches 3.0 10.0 NA NA 0.35
Active Displacement Inches 1.5 4.0 NA NA 0.35
Transverse Displacement Inches 1.0 4.0 NA NA 0.35
The probability that the seismic demands (D) placed on a component exceed the
capacity (C) conditioned on a chosen intensity measure (IM) can be assessed by a fragility
function. Assuming a lognormal distribution for the D and C, component fragility curves,
defined here as the probability of reaching or exceeding a specified damage state for a
130
ln( S / S )
P[ D C / IM] d c
(6.3)
d2/ IM c2
where, Sd is the median estimate of the demand as a function of the IM, Sc is the
median estimate of the capacity, d/IM is the dispersion of the demand conditioned on the
IM, c is the dispersion of the capacity, and (•) is the standard normal cumulative
computed through a convolution of a PSDM and a limit state model. A set of component
fragility curves computed in equation 6.3 has to be integrated into a system fragility (or
bridge fragility), which is facilitated through the development of joint probabilistic seismic
demand models (JPSDMs) (Nielson 2005, Padgett 2007, Ramanathan et al. 2015, Jeon et
al. 2015). The JPSDM recognizes the correlation between various components. If the
vector demands, Xi, placed on the n components of the system are expressed as
in the log-transformed space. The JSPDM is formulated in this space by assembling the
vector of means Y and the covariance matrix, Y . A Monte Carlo simulation (106 in
the current study) in which samples are drawn from both the demand and capacity models
is used to estimate the probability of demand exceeding the capacity value for each IM.
This procedure is repeated for the increasing value of the IM, and regression analysis is
used afterwards to estimate the lognormal parameters, median, and dispersion, which
characterize the bridge fragility. The system fragilities are helpful to measure the
dominating the overall system vulnerability. The system or bridge fragilities are useful for
calculate the probability of failure of each bridge component, and then combine them to
form the probability of failure of the system. Moreover, it is found that using the fragility
131
of any single bridge component to represent the overall vulnerability of the bridge would
6.3 shows a schematic of the fragility framework adopted in the current study.
The input for the fragility framework is presented in the following section.
132
6.2.1 Ground motion suite
peak ground accelerations to ensure the evaluation of a sufficient range of bridge responses.
The current study utilizes the ground motions from the NGA-2 database assembled by
Caltrans (2017). These motions were developed specifically for this project and consist of
320 ground motions. The ground motion details are given in Appendix A, and the response
spectra for the ground motions are presented in Figure 6.4. The work of Ramanathan (2012)
indicated that spectral acceleration at 1.0 sec (Sa-1.0s) is the optimal intensity measure for
the class of concrete box-girder bridges. Based on this observation, the current study
133
6.2.2 Material and geometric uncertainties and parameterized stochastic bridge models
Geometric and material uncertainties considered in the current study are discussed
in detail in Chapter 5. As mentioned before, most of the parameters were chosen based on
the plan review of more than 1,000 bridges, using in-house database obtained from
Caltrans. Having identified all key modeling assumptions and the uncertainty distribution
of the bridge models, the next step was to develop representative bridge models that could
capture the entire range of material and geometric uncertainties. Statistically significant yet
the range of parameters using Latin Hypercube Sampling technique, and were then paired
randomly with the selected suite of ground motions. 320 analytical bridge models were
generated, consistent with the number of ground motions. These were then paired randomly
to create the bridge model-ground motion pair. NLTHA was performed on each case and
the peak component demands are noted for each. The responses from the NLTHA were
used to generate the PSDMs. The PSDMs were then convolved with capacity models to
6.3 Fragility curves for multi-span continuous concrete single frame box-girder
bridges
The intent of this research is to generate fragility curves for multi-span continuous
relative vulnerabilities of various bridge groupings. Some nomenclatures were adopted for
this study and are detailed in Table 6.14. For example, S-E1-S22-R-D corresponds to
single-column bent (S) in design era Era11 (E1) with two spans (S22), rectangular cross-
134
One simple technique to evaluate the differences in fragility curves is to evaluate
the relative change in the median value of the fragility curves. An increase in the median
value means a less vulnerable structure, while a decrease in the median value indicates a
more vulnerable structure, and is illustrated in Figure 6.5. In Figure 6.5, bridge 2 is less
Fragility curves were generated for the 72 single frame box-girder bridge classes
and are presented in Table 6.15 - 6.18. These classifications cover more than 75% of the
California box-girder bridge inventory. Tables 6.15 – 6.18 also give the dispersion, , a
single value of dispersion characterizing the fragility across the four limit states. Appendix
B documents the fragility values in terms of PGA. The system and component fragility
curves for two bridge classes, S-E1-S22-C-D and S-E2-S22-C-S, are presented in Figure
6.6. It is seen from Figure 6.6 that the column is the most vulnerable component for bridges
135
with diaphragm abutments. In the case of bridges with seat abutments, columns and bearing
each contribute to the system vulnerability. The median and dispersion values for the
component fragility curves for the bridge classes are documented in Appendix C.
136
Table 6.15 – Fragility values for two span continuous concrete box-girder fragilities
with diaphragm abutments.
Design Bridge class BSST-0 BSST-0 BSST-0 BSST-0 *
era
S-E1-S22-C-D 0.12 0.60 0.30 0.61 0.70 0.63 1.06 0.62 0.61
S-E1-S22-R-D 0.17 0.63 0.41 0.64 0.82 0.65 1.18 0.66 0.65
T-E1-S22-C-D 0.08 0.63 0.20 0.66 0.47 0.71 0.72 0.70 0.68
Era 11
T-E1-S22-R-D 0.10 0.69 0.21 0.73 0.43 0.75 0.62 0.76 0.73
M-E1-S22-C-D 0.07 0.82 0.17 0.87 0.44 1.11 0.69 1.11 0.98
M-E1-S22-R-D 0.06 0.88 0.15 1.06 0.41 1.36 0.69 1.36 1.17
S-E2-S22-C-D 0.15 0.68 0.63 0.68 1.45 0.69 1.93 0.69 0.68
S-E2-S22-O-D 0.16 0.64 0.71 0.64 2.36 0.73 3.07 0.65 0.67
T-E2-S22-C-D 0.11 0.62 0.42 0.61 0.97 0.66 1.27 0.66 0.63
Era 22
T-E2-S22-O-D 0.15 0.58 0.55 0.57 1.17 0.57 1.47 0.57 0.57
M-E2-S22-C-D 0.10 0.61 0.33 0.60 0.67 0.74 0.86 0.74 0.68
M-E2-S22-O-D 0.10 0.61 0.33 0.60 0.67 0.74 0.86 0.74 0.68
S-E3-S22-C-D 0.15 0.68 0.63 0.68 1.91 0.70 2.86 0.72 0.69
S-E3-S22-O-D 0.16 0.64 0.71 0.64 3.10 0.67 4.39 0.63 0.65
T-E3-S22-C-D 0.11 0.62 0.41 0.61 1.27 0.66 1.84 0.66 0.64
Era 33
T-E3-S22-O-D 0.15 0.58 0.55 0.57 1.49 0.57 2.06 0.58 0.57
M-E3-S22-C-D 0.10 0.63 0.33 0.60 0.86 0.75 1.20 0.74 0.68
M-E3-S22-O-D 0.10 0.63 0.33 0.60 0.86 0.75 1.20 0.74 0.68
Table 6.16 – Fragility values for two span continuous concrete box-girder fragilities
with seat abutments.
Design Bridge class BSST-0 BSST-0 BSST-0 BSST-0 *
era
S-E1-S22-C-S 0.08 0.61 0.15 0.59 0.29 0.58 0.42 0.58 0.59
S-E1-S22-R-S 0.08 0.52 0.15 0.54 0.27 0.53 0.38 0.52 0.53
T-E1-S22-C-S 0.06 0.57 0.11 0.60 0.21 0.59 0.30 0.59 0.59
Era 11
T-E1-S22-R-S 0.08 0.57 0.14 0.57 0.25 0.57 0.35 0.55 0.57
M-E1-S22-C-S 0.04 0.63 0.08 0.62 0.18 0.60 0.26 0.60 0.61
M-E1-S22-R-S 0.09 0.58 0.15 0.58 0.27 0.58 0.37 0.58 0.58
S-E2-S22-C-S 0.11 0.58 0.53 0.57 0.99 0.65 1.32 0.65 0.61
S-E2-S22-O-S 0.11 0.51 0.53 0.49 1.14 0.55 1.58 0.55 0.52
T-E2-S22-C-S 0.08 0.51 0.38 0.49 0.77 0.55 1.08 0.56 0.53
Era 22
T-E2-S22-O-S 0.08 0.45 0.37 0.44 0.80 0.55 1.10 0.56 0.50
M-E2-S22-C-S 0.07 0.49 0.36 0.48 0.73 0.59 1.02 0.62 0.54
M-E2-S22-O-S 0.08 0.54 0.42 0.51 0.91 0.66 1.30 0.67 0.60
S-E3-S22-C-S 0.11 0.57 0.53 0.56 1.25 0.63 1.88 0.63 0.60
S-E3-S22-O-S 0.12 0.50 0.53 0.49 1.28 0.53 1.91 0.53 0.51
T-E3-S22-C-S 0.08 0.51 0.38 0.50 0.94 0.53 1.41 0.54 0.52
Era 33
T-E3-S22-O-S 0.08 0.45 0.37 0.43 0.91 0.52 1.32 0.50 0.48
M-E3-S22-C-S 0.07 0.49 0.36 0.49 0.89 0.58 1.35 0.58 0.53
M-E3-S22-O-S 0.08 0.54 0.42 0.52 1.07 0.60 1.64 0.61 0.56
137
Table 6.17 – Fragility values for multi-span continuous concrete box-girder
fragilities with diaphragm abutments.
Design Bridge class BSST-0 BSST-0 BSST-0 BSST-0 *
era
S-E1-S34-C-D 0.12 0.61 0.33 0.65 0.83 0.70 1.33 0.69 0.66
S-E1-S34-R-D 0.06 0.89 0.25 1.11 1.00 1.32 2.04 1.31 1.16
T-E1-S34-C-D 0.06 0.72 0.15 0.77 0.36 0.89 0.56 0.89 0.82
Era 11
T-E1-S34-R-D 0.07 0.75 0.17 0.79 0.41 0.91 0.66 0.92 0.84
M-E1-S34-C-D 0.01 1.04 0.04 0.98 0.19 1.13 0.40 1.13 1.07
M-E1-S34-R-D 0.01 1.04 0.04 0.98 0.19 1.13 0.40 1.13 1.07
S-E2-S34-C-D 0.15 0.56 0.59 0.54 1.34 0.71 1.79 0.71 0.63
S-E2-S34-O-D 0.19 0.67 0.93 0.66 3.04 0.93 4.26 0.87 0.78
T-E2-S34-C-D 0.09 0.62 0.39 0.59 0.82 0.89 1.12 0.89 0.75
Era 22
T-E2-S34-O-D 0.11 0.71 0.56 0.73 1.79 1.12 2.64 1.16 0.93
M-E2-S34-C-D 0.09 0.54 0.35 0.50 0.68 0.71 0.93 0.71 0.61
M-E2-S34-O-D 0.03 0.82 0.47 0.73 2.00 1.59 4.00 1.66 1.20
S-E3-S34-C-D 0.15 0.55 0.58 0.54 1.78 0.71 2.63 0.69 0.62
S-E3-S34-O-D 0.20 0.67 0.93 0.66 4.47 0.94 6.44 0.82 0.77
T-E3-S34-C-D 0.09 0.62 0.39 0.60 1.12 0.89 1.72 0.89 0.75
Era 33
T-E3-S34-O-D 0.11 0.73 0.56 0.73 2.59 1.12 4.34 1.12 0.93
M-E3-S34-C-D 0.08 0.54 0.35 0.50 0.92 0.70 1.40 0.71 0.61
M-E3-S34-O-D 0.03 0.86 0.47 0.74 3.74 1.47 9.02 1.50 1.14
138
S-E1-S22-C-D S-E2-S22-C-S
139
Figure 6.6 – System and component fragility curves for bridge classes S-E1-S22-C-D
and S-E2-S22-C-S.
The plot of the median value of bridge fragility curves for two bridge classes, S-E-
S22-C-D and T-E-S34-C-S is presented in Figure 6.7, and the trend is similar across the
different bridge classes. The following are the salient inferences noted from bridge
• In general, Era 11 bridges are more vulnerable than Era 22 and Era 33 bridges. This
• Era 22 bridges are more vulnerable than Era 33 bridges. The lower vulnerability of
Era 33 is due to the high ductility that is associated with Era 33 bridge columns.
• Across the design eras for a particular abutment type, it is seen that single column
bents are more vulnerable than the multi-column bents. This trend is consistent with
a previous study (Mangalathu et al. 2016a). Amongst the bridges with multi-
140
• In general, diaphragm abutment bridges are less vulnerable than seat abutment
bridges for all design eras. The lower vulnerability of the diaphragm abutment
a) 3.50
3.00 Era 11 Era 22 Era 33
BSST-i(i=0,1,2,3)
2.50
2.00
1.50
1.00
0.50
0.00
BSST-0 BSST-1 BSST-2 BSST-3
b) 2.00
1.80 Era 11 Era 22 Era33
BSST-i(i=0,1,2,3)
1.60
1.40
1.20
1.00
0.80
0.60
0.40
0.20
0.00
BSST-0 BSST-1 BSST-2 BSST-3
Figure 6.7 – Plot of median values of bridge classes across the design eras for a) S-
E(1/2/3)-S22-C-D, and b) T-E(1/2/3)-S34-C-S.
141
6.3.2 Trends based on spans
The variation of the median value of fragility curves for bridge classes with number
of spans is presented in Figure 6.8. The following inferences can be made from the data
• The relative vulnerability between the bridge classes with different spans depends on
the column cross-section, abutment type, design era, and number of columns per bent.
For example, it is seen in Figure 6.8 that the multi-span Era 22 bridge class with circular
its two-span counterpart S-E2-S22-C-D for all limit states. The percentage change in
median values between two-span and multi-span bridges are 1%, 7%, 8%, and 8% for
BSST-0, -1, -2, and -3 respectively. However, the trend is different for the Era 22 bridge
class with oblong cross-section resting on diaphragm abutments. In this case, multi-
span bridges are less vulnerable than two-span bridges, and the relative vulnerabilities
are 19%, 32%, 29%, and 39% for BSST-0, -1, -2, and -3 respectively.
• The differences in vulnerabilities for two-span and multi-span bridges underscore the
necessity to account for the number of spans in the generation of fragility curves, which
142
5.00
S-E2-S22-C-D S-E2-S34-C-D
4.00
BBST-i(i=0,1,2,3)
S-E2-S22-O-D S-E2-S34-O-D
3.00
2.00
1.00
0.00
BSST-0 BSST-1 BSST-2 BSST-3
Bridge system Damage States
Figure 6.8 – Plot of median values of bridge classes across the number of spans for
bridge classes S-E2-S(22/34)-C-D and S-E2-S(22/34)-O-D.
3.50
S-E1-S22-C-D S-E1-S22-C-S
3.00 S-E2-S22-C-D S-E2-S22-C-S
S-E3-S22-C-D S-E3-S22-C-S
BBST-i(i=0,1,2,3)
2.50
2.00
1.50
1.00
0.50
0.00
BSST-0 BSST-1 BSST-2 BSST-3
Bridge system Damage States
Figure 6.9 – Plot of median values of bridge classes across the type of abutments for
the selected bridge classes.
The following inferences can be deduced from the comparison of fragilities for
143
• For all design eras, diaphragm abutment bridges are less vulnerable than seat abutment
bridges. A general trend is shown in Figure 6.9 for selected bridge classes. The lower
engagement of the superstructure and the abutments in the load transfer mechanism.
• The columns are the most vulnerable component of bridges with diaphragm abutments.
However, in the case of seat abutment bridges, bearings as well as columns contribute
to the overall vulnerability. This highlights the need for adequate seat width in the case
• The vulnerability of bridges reduces with the evolution of column design philosophy.
The trend is the same irrespective of the type of cross-section and number of columns
per bent.
• HAZUS suggests the same fragility relationships for bridges with seat and diaphragm
abutments. The fragility curves generated in the current study outline the need to
account for the type of abutments. For example, the bridge class with seat abutment (S-
E3-S22-C-S) is 36%, 19%, 53% and 53% more vulnerable than the bridge class with
diaphragm abutment (S-E3-S22-C-D) for the limit states BSST-0, -1, -2, and -3,
respectively. Note that the only difference between the bridge class S-E3-S22-C-S and
cross-sections, while Eras 22 and 33 utilize circular and oblong cross-sections. The
144
following inferences are noted from the comparison of median fragilities for bridges
classes (Tables 6.15 – 6.18, and Figure 6.10) with various cross-sections:
5.00
T-E1-S34-C-D T-E1-S34-R-D
T-E2-S34-C-D T-E2-S34-O-D
4.00 T-E3-S34-C-D T-E3-S34-O-D
BBST-i(i=0,1,2,3)
3.00
2.00
1.00
0.00
BSST-0 BSST-1 BSST-2 BSST-3
Bridge system Damage States
Figure 6.10 – Plot of median values of bridge classes across various column cross-
sections for the selected bridge classes.
• In the case of bridges from Eras 22 and 33, columns with oblong cross-sections are less
vulnerable than columns with circular cross-sections. This trend is the same
• No general trend is observed in the case of Era 11 bridges. For example, bridges with
rectangular cross-sections performed better in the case of two-span bridges with single
column bent and multi-column bent bridges with diaphragm abutments. However, the
trend is reversed in the case of two-span bridges with two-column bents resting on
diaphragm abutments. Also, there is no trend observed in the case of bridges resting on
seat abutments.
145
6.3.5 Trends based on number of columns per bent
The general trend of the change in the median value of the fragility curves is
presented in. The following inferences can be drawn from the comparison of the median
• Multi-column bents are more vulnerable than bridges with single column bents. The
increased vulnerability of bridges with multi-column bents is mainly due to the bridge
pinned at the base, while single-column bents have a significant amount of rotational
• For a given design era and number of spans, multi-column bents with circular cross-
section are more vulnerable. For example, M-E2-S34-C-D is 34%, 194%, and 331%
more vulnerable than M-E2-S34-O-D for the limit states BSST -1, -2, and -3,
tremendous reduction in the vulnerability of bridges from Eras 22 and 33, compared
• Further, the differences in the median value of bridge classes with different numbers
of columns per bent underscore the necessity to capture the number of columns per
bent in the fragility curves. Such a classification is not currently available in HAZUS.
146
5.00
a) S-E2-S34-C-D
S-E2-S34-O-D
4.00
T-E2-S34-C-D
BBST-i(i=0,1,2,3)
3.00 T-E2-S34-O-D
M-E2-S34-C-D
2.00 M-E2-S34-O-D
1.00
0.00
BSST-0 BSST-1 BSST-2 BSST-3
Bridge system Damage States
b) 2.50 S-E2-S34-C-S
S-E2-S34-O-S
2.00 T-E2-S34-C-S
BBST-i(i=0,1,2,3)
T-E2-S34-O-S
1.50
M-E2-S34-C-S
1.00 M-E2-S34-O-S
0.50
0.00
BSST-0 BSST-1 BSST-2 BSST-3
Bridge system Damage States
Figure 6.11 – Plot of median values of bridge classes across the number of columns
per bent for the selected bridge classes with a) diaphragm abutments and b) seat
abutments.
bridges based on seismic design, span length, bent type, and span discontinuity. However,
147
the current study utilized a performance-based grouping strategy. HAZUS failed to
consider the variability of geometric and material attributes, which this study incorporated.
Despite the differences between the present study and HAZUS, Sa (1.0s) is adopted as the
intensity measure in both cases. Also the number of damage states characterizing the bridge
system vulnerability is similar in both studies. Table 6.19 presents the median values of
HAZUS fragility curves corresponding to slight (s), moderate (m), extensive (e), and
complete (c) damage states. HAZUS suggests a single value of dispersion (ds) across all
bridge classes and damage states. The equivalent bridge class notations between HAZUS
and the current study are presented in Table 6.19 to facilitate comparison. The comparison
of HAZUS and the present study’s selected bridge class fragilities are shown in Figures
By comparing the HAZUS fragilities with the fragilities generated in this study, the
148
• For the lower limit states BSST-0 and BSST-1, bridges are more vulnerable than is
currently predicted by HAZUS. This trend is the same across all of the bridge classes
• Single column bent seat abutment bridges in Era 11 are more vulnerable than is
currently predicted by HAZUS (Figure 6.12). For example, for the respective limit
states BSST -1, -2, and -3, bridges in the S-E1-S22-C-S class are 338%, 200%, 90%,
• For all bridge classes, HAZUS either over-estimates or under-estimates the fragilities
for higher limit states BSST-2 and BSST-3. The over-estimation or under-estimation
depends on the column cross-section, design era, number of spans, and number of
• HAZUS fragilities are unreliable for Era 11 two-span and multi-span bridges (Figure
6.13) resting on seat and diaphragm abutments for all the limit states. For example,
HAZUS overestimates the bridge fragilities by 650%, 543%, 340% and 330%,
compared to T-E1-S22-R-S for the limit states BSST -1, -2, and -3 respectively.
• Era 33 two-column bridges with diaphragm abutments are less vulnerable than is
predicted by HAZUS.
• The dispersions obtained in the current study are close to the HAZUS values, except
for classes of multi-column bent bridges, where the dispersions are higher than HAZUS
values.
• Based on the current study, it is reasonable to conclude that existing HAZUS bridge
149
Figure 6.12– Comparison of HAZUS and selected bridge class fragilities for a) Era
11 bridges with single column bent and b) Era 22 and Era 33 bridges with single
column bent.
150
Figure 6.13– Comparison of HAZUS and selected bridge class fragilities a) Era 11
two-span bridges with two- and multi-column bents and b) Era 22 and Era 33 two-
span bridges with single and multi-column bents.
151
6.5 Closure
This chapter presents the multi-phase framework for the generation of fragility
curves, and the fragility curves for selected bridge classes. The framework includes the
analysis, convolution with limit-state models, and generation of system fragility curves
through joint probabilistic seismic demand models (JPSDMs). The variations in the
material and geometric properties of the bridges were accounted for on the basis of the plan
uncertainties and nonlinear responses of various bridge components was created using
OpenSees, and each of the bridge models obtained from a Latin hypercube sampling (LHS)
was randomly paired with one of ground motions. The response data of each component,
monitored from dynamic analyses, was used to develop the associated probabilistic seismic
demand models (PSDMs). An important aspect presented in this chapter is the suggestion
columns.
Bridge component and system fragilities were generated for 72 bridge classes and
are presented in detail in this chapter. The generated fragility curves were compared with
the existing HAZUS fragility curves. The following are some of the significant findings
• The seismic vulnerability for all bridge classes reduced with evolutions in column
• Multi-column bents are more vulnerable than single column bents. The increased
vulnerability of multi-column bents is mainly due to the bridge width and modeling
assumptions.
152
• Across the various design eras, diaphragm abutment bridges are less vulnerable than
• Columns are the most vulnerable components in the case of bridges with diaphragm
• There is a wide disparity between the fragility curves generated in the present study
and the existing HAZUS fragility curves. Based on the current study, it is reasonable
to state that existing HAZUS bridge groupings and their associated fragility curves
153
CHAPTER 7 PARAMETERIZED FRAGILITY CURVES: LASSO
APPROACH
variables and level of treatment needed for these variables in the estimation of seismic
demand models and fragility curves. As seen in the previous chapters, the seismic fragility
of bridges has been expressed with one-dimensional (1-D) fragility curves developed with
parameter (IM). It is difficult to estimate the sensitivity of seismic demand models to input
conditioned only on IM. Also, as stated in Ghosh et al. (2013), single-parameter demand
models and fragility curves have some limitations: (1) the inability to account for the
earthquakes without extensive re-simulations for each new set of parameter combinations,
(2) the inability to explicitly address the effect of uncertainty parameters on fragility
curves, and (3) the lack of flexibility to incorporate field instrumentation data resulting
predictor variables has been gradually increased in the realm of seismic vulnerability and
loss estimation of bridges (Seo and Linzell, 2012; Dukes, 2013; Dukes et al. 2017 Ghosh
et al., 2013; Kameshwar and Padgett, 2014; Park and Towashiraporn 2014; Mangalathu et
al., 2015; Jeon et al., 2017; Mangalathu et al. 2017c). Assuming that the input variables are
(demand parameter) is constructed. Samples obtained from this demand model are
compared with those of the associated limit-state model to obtain the binary survival-
154
failure vector. This vector is used to perform a logistic regression analysis to determine the
regression coefficients and thus develop the multi-parameter fragility curve in the
component. This chapter (1) identifies the variables that exhibit strongest influences on the
seismic demand and seismic fragilities, (2) quantifies the relative impact of various sources
response surface models such as linear, stepwise, Lasso, Ridge, and elastic net in the
methodology that accounts for the effect of uncertain input variables in the seismic demand
models as well as seismic fragilities. Such a study (1) provides insight in quantifying
(2) eliminates the parameters which have a minimal influence on the seismic demand and
reduces unnecessary and exhaustive efforts in statistical sampling, (3) identifies the
parameters that can reduce the uncertainty in demand models and fragility curves by more
database), and (4) helps bridge owners, such as California Department of Transportation,
spend their resources judiciously (e.g. data collection, field investigations, censoring) on
The generation of the parameterized fragility curves for the bridge classes discussed
in Chapter 6 is beyond the scope of this thesis. It is noteworthy to mention that the
geometric, material, and structural uncertainties, ground motions, and fragility curves in
this chapter are used to demonstrate the approach and are not consistent with the earlier
chapters. Note that the purpose of this chapter is to suggest a methodology and demonstrate
its application. The approach is explained with a case study of two-span box girder bridges
in California. A brief review of various regression models, such as linear, stepwise, Lasso,
Ridge, and elastic net, is given in the next section. The efficiency of these models in
155
estimating the seismic demand is then compared using the mean square error (MSE) and
The following subsections describe the various regression models used for
estimation of seismic demand models. Five regression models such as linear regression,
stepwise regression, Ridge regression, Lasso regression, and elastic net regression are used
in this paper; this subsection describes their relevance to seismic demand modeling.
The linear regression (or least squares fitting) is the simplest and most commonly
applied form of regression technique and provides a solution to the problem of finding the
best-fitting straight line through a set of points. In the case of the seismic demand model
for bridges with input vector, X = (X1, X2,..., Xp) and a real-valued seismic demand (output
where j’s are the unknown parameters or coefficients and p is the number of input
parameters. With a training data set ( x1 , d1 ),..., ( xN , d N ) the most popular method for the
estimation of is to minimize the residual sum of squares (RSS, Equation 7.2). The training
156
The least square estimates of the parameter have the smallest variance among all
In the stepwise regression approach, only a subset of the input variable is retained
and the rest of the variables are eliminated by a selection criterion. Forward stepwise
regression is used in this paper. Forward stepwise regression starts with the intercept, and
subsequently adds the variables into the model that most improve the fit. The improvement
in fit is often based on the F statistic (Equation 7.3) in which the variables are sequentially
RSS ( ˆ ) RSS ( )
F (7.3)
RSS ( ) / ( N k 2)
The parameter estimate ˆ in Equation 7.3 is with k inputs and the estimate is with
the addition of a predictor. The F-ratio stopping rule doesn’t attempt to find the best model,
as the stopping rule provides only local control of the model search (Friedman et al. 2001).
The least square estimates often have low bias but suffer the drawback of large
variance (Hoerl and Kennard, 1970). The search for a biased estimator with smaller mean
square error (MSE) and significant reduction in variance led to the development of Ridge
regression (Tibshirani, 1996). Regression coefficients are shrunk by the Ridge regression
by imposing a penalty on their size; the Ridge regression minimizes a penalized residual
157
N p p
ˆridge argmin (di 0 xij j ) 2 j2 (7.4)
i1 i 1 j 1
The in Equation 7.4 corresponds to the shrinkage parameter, as the larger the
value of , the greater the shrinkage (towards zero) of the regression coefficients. The
inputs have to be standardized before applying the Ridge regression because the Ridge
solutions are not equivariant under the scaling of the input variables. Also, the intercept 0
(Tibshirani, 1996). Lasso regression minimizes the residual sum of squares, subjected to a
constraint based on the sum of absolute values of regression coefficients (Equation 7.5).
N p p
ˆlasso argmin (di 0 xij j )2 j (7.5)
i1 i 1 j 1
the Lasso penalty, Lasso regression does variable selection and shrinkage of regression
coefficients. In other words, the most significant variables are retained while the
insignificant variables are removed from the model. Bias is more controllable in Lasso
158
7.1.5 Elastic net
Zou and Hastie (2005) suggested another regularization and variable selection
N p
1 p p
ˆelastic argmin (di 0 xij j )2 j2 j (7.6)
i1 i 1 2 i 1 j 1
a bridge between Lasso regression and Ridge regression. It is particularly useful for
analyzing high dimensional data. In Ridge, Lasso, and elastic net, needs to be specified
by the analyst.
evaluation of seismic demand model will be discussed in Section 7.3. More detailed
descriptions of the various regression techniques can be found in Friedman et al. (2001).
Two-span box girder bridges are the most common type of highway bridge
inventory in California and account for more than 35% of the box girder bridge inventory.
Two-span bridges with seat and diaphragm abutments are selected for the case study in this
paper, and the selected bridges were designed and constructed prior to 1970. Figure
numerical modeling is carried out with the help of the finite element package OpenSees. A
159
Different sources of geometric, material, and system uncertainties are included in
this study. Table 7.1 shows the mean value (), standard deviation (), and the associated
probability distribution of various input variables used in the current study. Mangalathu et
al. (2016a) identified the input variables based on an extensive plan review and hence
mimic the California bridge inventory. The current study selects the suite of ground
motions developed by Baker et al. (2011), which was proposed as part of the PEER
Transportation Research Program. The suite comprises 120 pairs of broadband ground
motions and 40 pairs of near-fault ground motions. The entire suite of ground motions are
scaled by a factor of two (Ramanathan, 2012) to have sufficient response data of IMs higher
than the Palmdale spectrum (the highest probabilistic design hazard level in California),
and thus the expanded suite of 320 ground motions is used for the current study. The
spectral acceleration at 1.0 sec (Sa–1.0s) is adopted as the intensity as the IM in the current
The input variables are sampled across the range of parameters presented in Table
7.2 using Latin Hypercube Sampling technique to generate statistically significant yet
nominally identical three–dimensional bridge models. The variables are randomly paired
with the selected suite of ground motions. The two orthogonal components of the ground
motions are randomly assigned to the longitudinal and transverse direction of the bridge
axis. A set of NLTHAs (320 simulations) is performed for all bridge-ground motion pairs
to monitor the maximum response of various bridge components. The various demand
160
Figure 7.1 – Numerical modeling of various bridge components.
161
Table 7.1 – Uncertainty distribution considered in the bridge models.
Distribution
Parameter Units
Type μa σ*
Concrete compressive strength (fc) MPa Normal 29.03 3.59
Reinforcing steel yield strength (fy) MPa Lognormal 465.0 37.30
Span length (L) mm Lognormal 31775 8738
Deck width (Bd) mm Lognormal 9780 1980
Column height (H) mm Lognormal 6625 865
Abutment backwall height (Ha)
Diaphragm abutments
On piles mm Lognormal 3234 488
On spread footings mm Lognormal 2925 1056
Seat-type abutments
On piles mm Lognormal 2186 441
On spread footings mm Lognormal 2186 441
Abutments on piles - Lateral capacity/deck width (Kpa)
Diaphragm abutment N/mm Lognormal 1120 404
Seat-type abutment N/mm Lognormal 1498 540
Elastomeric bearing pad
Stiffness per deck width (Kb) N/mm/m Lognormal 908 327
Coefficient of friction for bearing pad (b) – Normal 0.30 0.10
Gap (g)
Longitudinal (btw. deck and abutment wall, l) mm Lognormal 23.5 12.5
Transverse (btw. deck and shear key, t) mm Lognormal 12.8 2.58
Mass factor (m) Uniform 1.25 0.007
Damping () Normal 0.045 0.0125
Acceleration for shear key capacity (as) g Lognormal 1.00 0.20
Longitudinal reinforcement ratio () (%) Uniform 2.25 0.52
Pile group – pile cap and piles
Translational stiffness (Kft)
Single column – 1% long. rebar N/mm Normal 297716 140101
Single column – 3% long. rebar N/mm Normal 245178 105076
Multi column – 1.5% long. rebar N/mm Normal 140101 105076
Rotational stiffness (Kfr)
Single column – 1% long. rebar N-m/rad Normal 4.5109 1.1109
Single column – 3% long. rebar N-m/rad Normal 6.810 9
1.1109
Multi column – 1.5% long. rebar N-m/rad Normal 0 0
Superstructure box type
Reinforced vs. cast-in-place prestressed concrete(BT) Bernoulli Equally split**
Abutment backfill type (sand vs. clay, ST) Bernoulli Equally split**
Ground motion direction (fault parallel)
Longitudinal vs. transverse (ED) Bernoulli Equally split**
Column diameter (1524 mm vs. 1828.8 mm, D) Bernoulli Equally split**
*µ and denotes the mean and standard deviation of the distribution. ** 160 simulations are carried out
by one type and the remaining 160 simulations are carried out by other type and are chosen randomly
162
Table 7.2 – Bridge component demand parameters.
Bridge component Demand parameter Abbreviation Units
Column Curvature ductility COL –
Abutment Passive abutment displacement ABP mm
Active abutment displacement ABA mm
Transverse abutment displacement ABT mm
Deck Displacement DEC mm
Bearing Superstructure unseating UST mm
displacement
Bearing deformation BRD mm
Foundation Translation FNT mm
Rotation FNR rad
As the variable of interest ranges over several orders of magnitude (Table 7.1), the
demand and input variable are transformed into the logarithmic space (Cornell et al. 2000;
Mangalathu et al. 2016b). Also, because Ridge, Lasso, and elastic net regression models
are sensitive to the scaling of input variables, the input variables are transformed (after the
logarithmic transformation) into standard space (zero mean and unit variance). It is
assumed that all of the input parameters are independent of each other and are therefore
non-correlated. As mentioned before, Ridge, Lasso, and elastic net models identify the
variables that have less influence on the regression model by penalizing the regression
coefficient associated with the variable to zero. The regression coefficient of the input
variable that has a minimal variance on the regression is penalized and the penalization
procedure depends on the type of regression formulation (Equation 7.4, 7.5 and 7.6). A
detailed description of the penalization procedure and algorithms can be found in Friedman
et al. (2001). In general, the larger the penalty applied, the greater the shrinkage of
regression coefficients (or setting the regression coefficient associated with the least
163
significant variables in the regression model to zero). However, an increase in penalization
beyond a certain point might lead to the removal of variables that have a significant
influence on the demand model. Hence, an investigation is carried out to find the optimal
penalty factor.
The performance or penalization of Ridge, Lasso, and elastic net regression models
depends on the shrinkage factor (Equation 7.4, 7.5 and 7.6). Thus, this study is carried
out to (1) understand the variation of MSE with and (2) determine the optimal value of
for the estimation of seismic demand models. The MSE for a vector D with n predictors
can be estimated as
1 n
MSE ( Di Di )2
n i 1
(7.7)
MSE is estimated in this study through tenfold cross validation: the data is fitted on nine-
tenths of the data, and the prediction error is computed for the remaining data. An in-depth
discussion on the cross-validation techniques can be found in Friedman et al. (2001). Note
that = 0 corresponds to the linear regression model, and in the case of = , all the
regression coefficients except intercept are shrunk to zero. The results of the investigation
of shrinkage factor are presented in Figure 7.2 and Figure 7.3. Figure 7.2 shows some
bridge components for the variation of MSE with , while Figure 7.3 shows the shrinkage
of the regression coefficients with . Figure 7.3 measures the number of regression
coefficients (or significant parameters) retained in the model with the increase in . The
following inferences obtained from Figure 7.2 and Figure 7.3 are summarized below.
164
• MSE decreases with until reaching an optimal value for Ridge, Lasso and elastic
• The MSE of Ridge, Lasso, and elastic net regressions decreases with the increase of
up to the optimal value, and thus the Ridge, Lasso, and elastic net regressions
• The optimal varies depending on the component demand parameter. For example,
in the case of diaphragm abutment bridges, the optimal for COL is 0.014 in the case
also increases the MSE of the demand model once is beyond the optimal .
• As the recommended values of (Friedman et al. 2001) are very low (in the order of
10-2 to 10-3), Ridge regression is not able to shrink the regression coefficients.
Although only selected component demand parameters are shown in Figure 7.2, similar
165
Figure 7.2 – Comparison of MSE with .
The optimal value of is adopted as 0.01 in the current study because it is found to
be the lower bound for all the optimal values in the estimation of seismic demand
parameters. However, such an estimate fails to identify many insignificant variables (or
fails to shrink the regression coefficient), and thus overestimates the shrinkage coefficients.
Nevertheless, the criteria adopted for the selection of is that the MSE of the models
should (1) be less than the linear regression, (2) shrink highly insignificant regression
coefficients, and (3) serve as a uniform value of for all the demand parameters. The
optimal choice of for each demand model is beyond the scope of this paper.
166
Figure 7.3 – Shrinkage of regression coefficients with .
The criteria adopted for evaluating the model-fitting procedure must be (1) able to
more accurately predict the future data and (2) a simple model with less number of
regression coefficients (Zou and Hastie, 2005). To evaluate the accuracy in predicting the
future data, the current database is split into two: training set and test set. 75% of the data
is assigned to the training set and is used to fit the model (Friedman et al. 2001). The
remaining 25% of the data is used as the test set to estimate the accuracy of the fitted model.
The assignment of data to the training set and test set is carried out randomly. The reason
to split the data into training set and test set is to avoid the over-fitting of data (Friedman
et al. 2001), which is a common problem if we use the entire data as training set. Table 7.3
167
shows the coefficients from five different regression models for COL with the training set
for the diaphragm abutment bridge. The estimated fit is used to check the error in the test
set. To achieve this goal, two types of error, MSE and absolute error (ABS), are selected
in this paper. The ABS for a vector Yˆ with n predictors can be estimated as
1 ˆ
ABS (Yi Yi ) (7.7)
n
Table 7.3 – Estimated coefficients and test error for COL for the bridge with
diaphragm abutments by various regression techniques
The standard error of the mean value of the MSE (std.MSE) and ABS (std.ABS) is
estimated as
standard deviation (MSE)
Std.MSE
n
standard deviation (ABS) (7.8)
Std.ABS
n
168
Table 7.3 also gives the MSE, Std.MSE, ABS, and Std.ABS value of the various
regression models. The comparison of MSE and ABS helps to demonstrate how well the
fit explains a given set of test data. The results from Table 7.3 show that (1) linear and
Ridge regressions identify all the input variables as significant; (2) stepwise regression is
the one which identifies the least number of significant parameters; (3) Lasso regression
model has the minimum of MSE, Std. MSE, ABS, and Std. ABS, and thus is the best fit
amongst the selected regression models; (4) among Lasso, Ridge, and elastic net, Lasso
regression identifies more insignificant parameters; (5) these three regression methods
identify that fy, Kft, and BT have less effect on the seismic demand model for COL, and
thus it can be deduced that these input parameters are the least significant parameters; (6)
all the regression methods identify Sa–1.0s as the most significant parameter; and (7)
standard error for all of the models are fairly similar. MSE and ABS are compared for the
different regression models for the various demand parameters and are plotted in Figure
7.4. Lasso regression is the one having the least MSE and ABS for all the demand models.
The MSE and ABS associated with COL, UST, and BRD are low for all of the regression
models, which shows the good predictive capability. However, the MSE and ABS for ABA
and ABP are higher for all the models. Previous studies (Ramanathan 2012; Mangalathu et
al. 2016a) pointed out that columns and bearings are the components that govern the system
fragility, and thus the adopted models are good enough to capture the seismic demand of
COL and UST. Note that the results shown in Table 7.3and Figure 7.4 are for = 0.01 and
the results can be significantly improved by selecting a better optimal value for . However,
the selected yields good results and is able to remove insignificant parameters from the
regression model. In addition, stepwise regression might result in the data over-fitting and
instability (Vidakovic 2011), so Lasso regression is adopted as the regression model for
further part of this study. Lasso regression leads to low values of MSE and ABS and has
169
Figure 7.4 – Radar plot depicting the comparison of accuracy of fit obtained from
the various regression models.
The regression coefficients from Lasso regression are also a measure of sensitivity
of input parameters to the seismic demand model or fragilities because the input parameters
170
are converted to a standard space (after a logarithmic transformation of the data) in Lasso
regression. A positive regression coefficient for a particular variable indicates that the
seismic demand increases with a positive increase in the variable, and a negative regression
coefficient shows that an increase in that variable reduces the seismic demand. Figure 7.5
shows the sensitivity of demand models for bridges with diaphragm and seat abutments. It
is noted from Figure 7.5 that the significant parameters for both bridges tend to vary from
component to component. In general, the IM (here, Sa–1.0s) has the greatest influence on the
demand model for all the demand parameters. Span length (L) is the second most sensitive
parameter for all the demand parameters. Soil type has a significant influence on the ABA
and ABP in the case of the diaphragm abutment bridges. The influence of soil type is due
to the fact that diaphragm abutments are stiffer than the adjacent bent and attract a large
portion of the seismic demand. Similar conclusions are also noted in previous studies
reinforcement ratio () and the column diameter (D) for the two case-study bridges. L and
Concrete strength (fc), steel strength (fy), damping ratio (), superstructure box type
(BT), earthquake direction (ED), abutment height (Ha), acceleration for shear key capacity
(as), gap between the deck and shear key (t), and coefficient of bearing (µb) all have a
minimal impact on all the seismic demand models. In addition, Jeon et al. (2017) used a
significant parameters affecting the seismic response of curved concrete box girder bridges.
Their work concluded that fc, fy, ED, µb, Ha, and t are the least significant parameters on
the seismic demand model of the bridges. The current sensitivity study shows that other
variables, except for IM, also have a significant influence on the seismic demand. The
demand model conditioned only on the IM, as used in traditional fragility analysis, might
not lead to a realistic estimation of the seismic demand and the fragility curves. Hence, a
171
fragility methodology accounting for the influence of significant parameters in the demand
Recently, a number of studies (Seo and Linzell, 2012; Dukes, 2013; Ghosh et al.,
2013; Kameshwar and Padgett, 2014; Park and Towashiraporn 2014; Jeon et al., 2015;
Mangalathu et al., 2015) generated parameterized component and system fragility curves
172
suggested in this paper. Unlike the previous studies, the proposed approach helps to
identify the relative impact of the various input parameters on the seismic demand as well
as fragilities. Also, the proposed approach removes the less significant variable from the
generation of seismic demand model and the seismic fragilities without much
computational effort. The outline of the proposed approach is given below, for uncertain
Step 1: Evaluate the linear regression coefficients (li) by performing Lasso regression
analysis for each component (ki, i = 1,…,m) with the input parameters (x1,…,xn,
Step 2: Generate a large number of demand estimates (N, 1 million in this study) for each
distribution.
Step 3: Generate N capacity values for a specific damage state for each bridge component
Step 4: Obtain the binary survive-failure (N 1) vector by comparing the capacity values
Step 5: Conduct a Lasso logistic regression on the survive-failure vector to determine the
nl
k ,0 k ,IM ln(IM) k j ln(x j )
j 1
e
PFk|IM, x1 , x2 ,.. xn nl
, nl n (7.9)
l
k,0 k,IM ln(IM) k j ln(x j )
1 e j 1
173
where k,0, k,sa, and k,j’s (j = 1,…,nl) are the Lasso logistic regression coefficients
of the kth bridge component. This step helps to identify the sensitivity of bridge
Step 6: Assuming that the bridge failure is a series system (the system fails if one or more
components fail), estimate the binary survive-failure vector, and conduct a Lasso
logistic regression to obtain the system failure. This step helps to identify the
ns
SYS ,0 SYS ,IM ln(IM) SYS , j ln(x j )
j 1
e
PFSYS |Sa , x1 , x2 ,.. xn ns
, ns n (7.10)
SYS , j ln(x j )
s
SYS ,0 SYS ,IM ln(IM)
1 e j 1
where SYS,0, SYS,sa, and SYS,j’s (j = 1,…, ns) are the Lasso logistic regression
Step 7: For a particular bridge with significant input parameters, x1 ,..., x ns , the classical one-
ns
SYS ,0 SYS ,IM ln(IM) SYS , j ln( x j )
j 1
e
PFSYS |IM ... ns
f ( x1 )... f ( xns ) dx1 ...dxns
x1 x2 xns SYS ,0 SYS ,IM ln(IM) SYS , j ln( x j )
1 e j 1
(7.11)
where f(x1),…, f(x9) are the probability density parameters for parameters,
x1 ,..., xns .
The limit state models for the various bridge components are given in Table 7.4,
and are consistent with the limit states presented in Chapter 6. The limit states were derived
174
in such a way as to align with the California Department of Transportation (Caltrans)
design and operational experience. This will facilitate Caltrans’ evaluation of repair–
and fragility curves are generated for the bridges with diaphragm and seat abutments for
various limit states. Figure 7.6 shows the single parameter fragility curves (conditioned on
Sa-1.0s) for the diaphragm and seat abutments for the moderate damage state. It is clear from
Figure 7.6 that the bridge fragility is mostly dominated by column fragilities for the
moderate damage state. ABT and BRD, respectively, are the second most vulnerable
components in the case of diaphragm abutment bridges and seat abutment bridges.
Although not shown here, similar conclusions are observed for other limit states. Interested
readers are directed to previous studies (Mangalathu et al. 2015, Jeon et al. 2017) for the
comparison of the fragility curves through the multi-parameter demand model with the
175
Figure 7.6 – System and component fragility curves for moderate damage state: a)
diaphragm abutment bridge, b) seat abutment bridge.
As mentioned before, the proposed approach also helps identify the sensitivity of
the fragility curves to the input parameters. Figures 7.7 and 7.8 show the sensitivity of the
fragility curves to the input parameters for the diaphragm and seat abutment bridges,
respectively. The uncertainty in IM dominated over all the other uncertainties in the
fragility curves. In the case-study bridges where the fragility of COL determines the bridge
fragility, Sa–1.0s, L, , D, Bd, and m are the most sensitive parameters for the diaphragm
abutment bridges for system fragility, while Sa–1.0s, L, , D, Bd, m, and l are the sensitive
parameters in the case of seat abutment bridges. fy, , ED, and Kfr are the least significant
parameters for all the components for all the damage states. Another advantage of the
proposed approach is the ability to identify the sensitivity of input variables based on the
limit state under consideration. For example, the sensitivity of IM is higher for ABP at the
moderate damage state when compared to the slight damage state in the case of diaphragm
coefficients.
176
Figure 7.7 – Sensitivity of fragility curves to input parameters for diaphragm
abutment bridge for various limit states
177
Figure 7.8 – Sensitivity of fragility curves to input parameters for seat abutment
bridge for various limit states
178
7.6 Conclusions
This chapter identifies the relative impact of various uncertain input parameters on
the seismic response of various bridge components. The efficiency of various regression
models such as linear, stepwise, Ridge, Lasso, and elastic net in the generation of seismic
demand models are evaluated in the initial part of the paper. The comparison is carried out
for two-span box girder bridges with seat and diaphragm abutments. Nonlinear time history
analysis (NLTHA) is carried out for the bridge models accounting for the material,
ductility demand, abutment displacements in the passive, active, and transverse directions,
displacement, and elastomeric bearing displacement are recorded for each NLTHA. Multi-
parameter demand models are generated for each demand parameter by using 75% of the
data from the NLTHA. The efficiency of the regression methods are compared in terms of
the mean square error (MSE) and absolute error (ABS) in predicting the remaining 25% of
the NLTHA data. It is observed that the Lasso regression model is the most effective with
regard to lowest MSE and ABS in predicting the data. Also, the Lasso regression model is
able to remove insignificant variables from the demand model. As the Lasso regression
coefficients are a measure of the sensitivity of the input variables, the results of the Lasso
regression is used to identify the significance of the input variables in the estimation of
seismic demand models for various bridge components. Ground motion intensity measure
(IM, here, spectral acceleration at 1 sec) and span length (L) are identified as the most
sensitive variables in the demand models for various demand parameters. In general, the
steel strength (fy), damping ratio (), superstructure box type (BT), earthquake direction
(ED), abutment height (Ha), acceleration for shear key capacity (as), gap between the deck
and shear key (t), and coefficient of bearing (µb) all have a minimal impact on all the
179
The sensitivity results reveal that the generation of demand models without
estimates of the demand models. To include the effect of significant uncertain variables in
methodology using Lasso regression is suggested in this paper. The proposed fragility
approach helps to identify the relative impact of the various input parameters on the
demand as well as fragility curves of various structural components. Hence, the proposed
approach provides additional perspectives for the decision makers or the owners in
sensitivity study on fragility curves indicates that IM, L, reinforcement ratio (), column
diameter (D), deck width (Bd), mass factor (m), and gap between the deck and shear key
(l) are the parameters significantly affecting on the bridge fragilities. In general, the
component as well as system fragilities are minimally affected by fy, , ED, and foundation
rotational stiffness (Kfr). The finding of the sensitivity study is helpful in evaluating the
level of uncertainty treatment required for each variable. Depending on the user
requirement and whether the intention is to have system or component vulnerability, the
Although the findings observed from this study are based on the case studies of
two-span concrete box girder bridges in California, the methodology is relevant and
challenge in regional risk assessment, the proposed approach helps to identify whether the
identification of level of uncertainty treatment required for a particular variable also helps
bridge owners spend their resources judiciously and develop a more reliable database of
180
CHAPTER 8 CONCLUSIONS AND FUTURE WORK
for each individual structure in a specific region. Also the generation of structure-specific
fragility curves is not warranted as some structures have similar performance or fragilities.
The grouping of structures is carried out in most cases based on engineering judgment and
critical review of HAZUS based on recent research revealed many drawbacks and showed
that the HAZUS bridge grouping and associated fragilities leads to an unrealistic estimation
of the seismic demand and the associated losses. These limitations in the traditional
techniques. The performance based grouping leads to more reliable sub-classes of bridges
relative to the traditional subjective lumping of bridges. The current study explores various
grouping of structures of similar performance. Based on the insights from the comparison,
attributes such as column cross-section, design era, number of spans, abutment type, pier
type and span continuity. Although the current study groups the box-girder bridge classes
181
in California, the proposed grouping approach can also be applied to other regions by fine-
tuning the grouping based on the evolution in seismic design philosophy and other
A major task in the current research was to understand and characterize the
California bridge inventory. California is a state with a high seismic hazard, a history of
damaging earthquakes, and has close to 29,000 bridges which vary in age based on their
construction. Bridge plans pertinent to various design eras and structural configurations
were reviewed in detail, and descriptive statistics were calculated for the material,
characterization helps to make the fragility models applicable to a wide geographic area.
More than 1000 bridge plans were reviewed for this process with the help of California
uncertainties are created in OpenSees. The numerical models incorporate a high degree of
detail with respect to the component modeling strategies and their ability to capture the
damage due to the seismic demand. The input variables are sampled across the range of
significant yet nominally identical three–dimensional bridge models. The variables are
randomly paired with the selected suite of ground motions. The two orthogonal
components of the ground motions are randomly assigned to the longitudinal and
transverse direction of the bridge axis. A set of NLTHAs (320 simulations) is performed
for all bridge-ground motion pairs to monitor the maximum response of various bridge
components. Various demand parameters considered in the current study are curvature
182
ductility, abutment displacements in the passive, active and transverse direction,
models are convolved with the capacity models to generate the component fragility curves.
A significant contribution of the present study is to suggest the capacity limit states (CCLS)
for columns based on extensive experimental review. Pertinent to various design eras,
literature review was carried out to collect the experimental data for bridge columns and
statistical analysis was carried out to suggest the CCLS of bridge columns. Such an exercise
helps to develop a new generation of more accurate and useful bridge fragility models for
incorporation into the ShakeCast earthquake alerting system developed by the California
in California. System fragility curves are generated afterwards using Monte Carlo
simulations and joint probabilistic seismic demand models (JPSDMs) incorporating the
Bridge component and system fragilities are generated for 72 bridge classes and the
selected bridge classes cover more than 75% of the California box-girder bridge inventory.
The following are some of the notable findings from the fragility analysis:
• The seismic vulnerability of all the bridge classes reduced with the evolution in
• Multi-column bents are more vulnerable than the single column bents. The increased
vulnerability of the multi-column bents is mainly due to the bridge width and
modeling assumptions.
• Across the various design eras, diaphragm abutment bridges are less vulnerable than
183
• Columns are the most vulnerable component in the case of bridges with diaphragm
abutments. However, in the case of seat abutment bridges, bearings also contribute
• Comparison with HAZUS fragilities revealed the wide disparity between the fragility
curves generated in the present study and the HAZUS fragilities. Based on current
curves through machine learning techniques. The framework includes the selection of a
demand models using the Lasso regression method, and development of parameterized
models are used (1) to produce bridge-specific (one-dimensional) fragility curves when the
fragility curves using a Monte Carlo integration. The advantage of the Lasso regression
model is that it is able to remove insignificant variables from the demand model. As the
Lasso regression coefficients are a measure of the sensitivity of the input variables, the
results of the Lasso regression are used to identify the significance of the input variables in
the estimation of seismic demand model for various bridge components. Ground motion
intensity measure (IM, here, spectral acceleration at 1 sec) and span length (L) are identified
as the most sensitive variables in the demand models for various demand parameters. In
general, the steel strength (fy), damping ratio (), superstructure box type (BT), earthquake
184
direction(ED), abutment height (Ha), acceleration for shear key capacity (as), gap between
the deck and shear key (t), and coefficient of bearing (µb) have a minimal impact on all
the seismic demand models. The proposed fragility approach helps to identify the relative
impact of the various input parameters on the demand as well as fragility curves of various
structural components. Hence, the proposed approach provides additional perspectives for
the decision makers or the owners in prioritizing their resources in the data base
This study presents a performance based grouping methodology to group the box-
girder bridge classes in California and generated component and system fragility curves
for single frame multi-span box girder bridges in California. This resulted in a significant
judgment and prior experience. This work (1) presents an overview of various statistical
and Kruskal Wallis (KW) test for grouping the bridges of similar performance; (2)
compares the groupings that emerge from the various grouping techniques; and (3)
identifies the method that has more statistical power in creating bridge sub-classes of
responses of bridge classes obtained from the non-linear time history analysis of
bridges.
185
• A simple performance based grouping methodology based on ANOVA is suggested in
this work. The ANOVA-based grouping provides insight into the potential bridge
attributes which significantly affect the seismic response and fragility curves, which to
date has not been assessed thoroughly. The method can be used to group bridge classes
depending upon the user requirement, i.e. whether the user would like to assess the
reliable estimation of the seismic vulnerability of various bridge classes. The proposed
method helps to save considerable computational effort and simulation compared to the
• Extensive details are provided regarding the variability in the material, structural and
• Column capacity limit states are suggested based on the literature review of
experimental studies on bridge columns. The proposed limit states are more reliable,
and help to develop a new generation of more accurate and useful bridge fragility
• Fragility curves and relative vulnerabilities are evaluated for the 72 box-girder bridge
classes in California. The selected bridge classes cover more than 75% of the California
box-girder bridge inventory. The generate fragility curves underscore the necessity to
• The study suggests a methodology to evaluate the sensitivity of demand models and
fragility curves to the uncertain input parameters. To include the effect of significant
uncertain variables in the generation of demand models and fragility curves, a multi-
186
parameter fragility methodology using Lasso regression is suggested in this work. The
proposed approach provides additional perspectives for the decision makers or the
monitoring.
Potential areas in which this work can be extended through additional research include the
following:
potential. Also, the random and non-liner behavior of soil can significantly affect
the behavior of bridges. Further studies are needed to explore the effect of
fragility curves.
• The performance based strategy should be extended to other bridge class such as I-
• The study focused on the seismic risk assessment due to main-shock ground
motions, and doesn’t account for aging and deterioration mechanism. The effect of
investigated further.
• The current study is limited to machine learning techniques such as Lasso, Ridge
and Elastic net. Future work should investigate other machine learning techniques
such as Random Forest, Support Vector Machine, to name a few. Also the
187
application of machine learning for fragility curves is limited to two-span box-
girder bridges with circular columns in Era 11. Future work can focus on expanding
the use of the method and tool to other bridge types common in California.
• The current study has adopted relevant capacity models for the bridge type and
design era of interest. In theory, the subgrouping of bridges presented in this paper
may also have slight variations in component capacities. However, such refined
capacity estimates that are attribute-dependent for a range of damage states and
components are still rather lacking and typically require extensive additional
experimental testing for model building or validation. As a result, the current study
has adopted capacity estimates appropriate for the bridge class and era, where
within the class. Further studies are needed for the development of refined capacity
models per damage state and component associated with variations in bridge
general capacity model per component, overall bridge class, and design era.
188
APPENDIX A. EARTHQUAKE RECORDS USED FOR FRAGILITY ANALYSIS
Aria
PEER
Scal Sa Vs30 s
Earth Record Rjb Rrup
e (1.0, (m/sec Inten Earthquake Name Year Station Name Magn Mechani
quake Sequence (km) (km)
Fact g) ) sity itude sm
no: Number
or (m/s
ec)
strike 91.2
1 0007 1.18 0.04 219.31 0 "Northwest Calif-02" 1941 "Ferndale City Hall" 6.6 slip 91.15 2
strike 56.8
2 0009 1.17 0.06 213.44 0.1 "Borrego" 1942 "El Centro Array #9" 6.5 slip 56.88 8
strike 27.0
3 0020 1.10 0.35 219.31 0.5 "Northern Calif-03" 1954 "Ferndale City Hall" 6.5 slip 26.72 2
4 0051 1.02 0.05 280.56 0 "San Fernando" 1971 "2516 Via Tejon PV" 6.61 Reverse 55.2 55.2
61.7
5 0056 0.90 0.03 235.00 0.1 "San Fernando" 1971 "Carbon Canyon Dam" 6.61 Reverse 61.79 9
"Cedar Springs 92.5
6 0058 0.87 0.02 477.22 0 "San Fernando" 1971 Pumphouse" 6.61 Reverse 92.25 9
"Gormon - Oso Pump 46.7
7 0065 1.20 0.07 308.35 0.1 "San Fernando" 1971 Plant" 6.61 Reverse 43.95 8
"LA - Hollywood Stor 22.7
8 0068 0.92 0.16 316.46 0.7 "San Fernando" 1971 FF" 6.61 Reverse 22.77 7
9 0070 1.12 0.37 425.34 0.3 "San Fernando" 1971 "Lake Hughes #1" 6.61 Reverse 22.23 27.4
28.9
10 0078 1.04 0.14 452.86 0.3 "San Fernando" 1971 "Palmdale Fire Station" 6.61 Reverse 24.16 9
"Wheeler Ridge - 70.2
11 0092 0.81 0.01 347.67 0 "San Fernando" 1971 Ground" 6.61 Reverse 68.38 3
12 0122 0.81 0.11 249.28 0.1 "Friuli Italy-01" 1976 "Codroipo" 6.5 Reverse 33.32 33.4
13 0126 2.11 1.33 259.59 5.7 "Gazli USSR" 1976 "Karakyr" 6.8 Reverse 3.92 5.46
14 0126 1.03 0.65 259.59 5.7 "Gazli USSR" 1976 "Karakyr" 6.8 Reverse 3.92 5.46
189
strike
15 0160 2.27 1.01 223.03 6.1 "Imperial Valley-06" 1979 "Bonds Corner" 6.53 slip 0.44 2.66
strike
16 0160 1.11 0.49 223.03 6.1 "Imperial Valley-06" 1979 "Bonds Corner" 6.53 slip 0.44 2.66
strike 10.4
17 0161 1.01 0.26 208.71 0.4 "Imperial Valley-06" 1979 "Brawley Airport" 6.53 slip 8.54 2
strike 10.4
18 0162 0.92 0.15 231.23 0.9 "Imperial Valley-06" 1979 "Calexico Fire Station" 6.53 slip 10.45 5
strike 21.6
19 0172 1.04 0.08 237.33 0.3 "Imperial Valley-06" 1979 "El Centro Array #1" 6.53 slip 19.76 8
strike 12.5
20 0174 2.45 0.58 196.25 2 "Imperial Valley-06" 1979 "El Centro Array #11" 6.53 slip 12.56 6
strike 12.5
21 0174 0.91 0.22 196.25 2 "Imperial Valley-06" 1979 "El Centro Array #11" 6.53 slip 12.56 6
strike 17.9
22 0175 0.91 0.16 196.88 0.4 "Imperial Valley-06" 1979 "El Centro Array #12" 6.53 slip 17.94 4
strike
23 0179 2.13 1.14 208.91 1.4 "Imperial Valley-06" 1979 "El Centro Array #4" 6.53 slip 4.9 7.05
strike
24 0179 0.86 0.46 208.91 1.4 "Imperial Valley-06" 1979 "El Centro Array #4" 6.53 slip 4.9 7.05
strike
25 0180 2.25 1.33 205.63 1.7 "Imperial Valley-06" 1979 "El Centro Array #5" 6.53 slip 1.76 3.95
strike
26 0180 1.09 0.65 205.63 1.7 "Imperial Valley-06" 1979 "El Centro Array #5" 6.53 slip 1.76 3.95
strike
27 0181 2.37 1.15 203.22 1.8 "Imperial Valley-06" 1979 "El Centro Array #6" 6.53 slip 0 1.35
strike
28 0181 1.05 0.51 203.22 1.8 "Imperial Valley-06" 1979 "El Centro Array #6" 6.53 slip 0 1.35
strike
29 0182 2.27 1.53 210.51 1.7 "Imperial Valley-06" 1979 "El Centro Array #7" 6.53 slip 0.56 0.56
strike
30 0182 0.98 0.66 210.51 1.7 "Imperial Valley-06" 1979 "El Centro Array #7" 6.53 slip 0.56 0.56
strike
31 0183 2.24 0.78 206.08 1.6 "Imperial Valley-06" 1979 "El Centro Array #8" 6.53 slip 3.86 3.86
strike
32 0183 0.90 0.31 206.08 1.6 "Imperial Valley-06" 1979 "El Centro Array #8" 6.53 slip 3.86 3.86
190
"El Centro Differential strike
33 0184 1.85 0.79 202.26 2.1 "Imperial Valley-06" 1979 Array" 6.53 slip 5.09 5.09
"El Centro Differential strike
34 0184 1.02 0.44 202.26 2.1 "Imperial Valley-06" 1979 Array" 6.53 slip 5.09 5.09
strike 30.3
35 0188 0.96 0.04 316.64 0.1 "Imperial Valley-06" 1979 "Plaster City" 6.53 slip 30.33 3
strike 31.9
36 0191 0.88 0.06 242.05 0.3 "Imperial Valley-06" 1979 "Victoria" 6.53 slip 31.92 2
37 0285 0.99 0.27 649.67 0.4 "Irpinia Italy-01" 1980 "Bagnoli Irpinio" 6.9 Normal 8.14 8.18
46.2
38 0287 0.86 0.04 356.39 0 "Irpinia Italy-01" 1980 "Bovino" 6.9 Normal 44.62 5
22.5
39 0288 1.00 0.10 561.04 0.5 "Irpinia Italy-01" 1980 "Brienza" 6.9 Normal 22.54 6
53.1
40 0294 0.87 0.05 496.46 0 "Irpinia Italy-01" 1980 "Tricarico" 6.9 Normal 51.74 6
"Taiwan 92.0
41 0427 1.03 0.02 671.52 0 SMART1(25)" 1983 "SMART1 E02" 6.5 Reverse 91.54 4
"Taiwan 97.6
42 0432 1.01 0.05 267.67 0 SMART1(25)" 1983 "SMART1 O01" 6.5 Reverse 97.16 3
43 0436 1.01 0.02 279.97 0 "Borah Peak ID-01" 1983 "CPP-601" 6.88 Normal 82.6 82.6
"TRA-642 ETR Reactor 79.5
44 0440 0.91 0.01 324.20 0 "Borah Peak ID-01" 1983 Bldg(Bsmt)" 6.88 Normal 79.59 9
"TRA-670 ATR Reactor
45 0441 1.07 0.02 324.20 0 "Borah Peak ID-01" 1983 Bldg(Bsmt)" 6.88 Normal 80 80
16.0
46 0587 0.99 0.21 551.30 0.7 "New Zealand-02" 1987 "Matahina Dam" 6.6 Normal 16.09 9
"El Centro Imp. Co. strike
47 0721 2.27 0.66 192.05 1.1 "Superstition Hills-02" 1987 Cent" 6.54 slip 18.2 18.2
"El Centro Imp. Co. strike
48 0721 0.95 0.28 192.05 1.1 "Superstition Hills-02" 1987 Cent" 6.54 slip 18.2 18.2
strike
49 0723 2.22 1.60 348.69 3.7 "Superstition Hills-02" 1987 "Parachute Test Site" 6.54 slip 0.95 0.95
strike
50 0723 1.08 0.78 348.69 3.7 "Superstition Hills-02" 1987 "Parachute Test Site" 6.54 slip 0.95 0.95
strike 22.2
51 0724 1.06 0.16 316.64 0.6 "Superstition Hills-02" 1987 "Plaster City" 6.54 slip 22.25 5
191
"Salton Sea Wildlife strike 25.8
52 0726 1.08 0.19 191.14 0.4 "Superstition Hills-02" 1987 Refuge" 6.54 slip 25.88 8
Reverse 23.9
53 0730 1.07 0.32 343.53 0.3 "Spitak Armenia" 1988 "Gukasian" 6.77 Oblique 23.99 9
Reverse 24.5
54 0737 0.95 0.16 239.69 0.5 "Loma Prieta" 1989 "Agnews State Hospital" 6.93 Oblique 24.27 7
"Anderson Dam Reverse 20.2
55 0739 0.91 0.16 488.77 0.8 "Loma Prieta" 1989 (Downstream)" 6.93 Oblique 19.9 6
Reverse 10.7
56 0741 2.30 1.23 476.54 5.4 "Loma Prieta" 1989 "BRAN" 6.93 Oblique 3.85 2
Reverse 10.7
57 0741 0.85 0.46 476.54 5.4 "Loma Prieta" 1989 "BRAN" 6.93 Oblique 3.85 2
"Bear Valley #14 Upper Reverse 71.3
58 0745 0.90 0.05 422.79 0.2 "Loma Prieta" 1989 Butts Rn" 6.93 Oblique 71.28 9
"Bear Valley #7 Reverse 69.3
59 0747 0.81 0.03 509.87 0 "Loma Prieta" 1989 Pinnacles" 6.93 Oblique 68.22 8
Reverse 44.1
60 0748 0.99 0.14 627.59 0.2 "Loma Prieta" 1989 "Belmont - Envirotech" 6.93 Oblique 43.94 1
Reverse
61 0753 2.47 1.24 462.24 3.2 "Loma Prieta" 1989 "Corralitos" 6.93 Oblique 0.16 3.85
Reverse
62 0753 1.20 0.61 462.24 3.2 "Loma Prieta" 1989 "Corralitos" 6.93 Oblique 0.16 3.85
Reverse 10.9
63 0764 1.06 0.39 308.55 0.7 "Loma Prieta" 1989 "Gilroy - Historic Bldg." 6.93 Oblique 10.27 7
Reverse 12.8
64 0767 2.11 0.67 349.85 2.1 "Loma Prieta" 1989 "Gilroy Array #3" 6.93 Oblique 12.23 2
Reverse 12.8
65 0767 1.09 0.35 349.85 2.1 "Loma Prieta" 1989 "Gilroy Array #3" 6.93 Oblique 12.23 2
"Hollister - South and Reverse 27.9
66 0776 1.77 1.26 282.14 2.2 "Loma Prieta" 1989 Pine" 6.93 Oblique 27.67 3
"Hollister - South and Reverse 27.9
67 0776 1.04 0.74 282.14 2.2 "Loma Prieta" 1989 Pine" 6.93 Oblique 27.67 3
Reverse
68 0779 1.58 1.19 594.83 7.9 "Loma Prieta" 1989 "LGPC" 6.93 Oblique 0 3.88
Reverse
69 0779 1.08 0.82 594.83 7.9 "Loma Prieta" 1989 "LGPC" 6.93 Oblique 0 3.88
192
"Salinas - John and Reverse 32.7
70 0800 1.00 0.10 279.56 0.2 "Loma Prieta" 1989 Work" 6.93 Oblique 28.66 8
"Saratoga - W Valley Reverse
71 0803 2.27 1.37 347.90 1.3 "Loma Prieta" 1989 Coll." 6.93 Oblique 8.48 9.31
"Saratoga - W Valley Reverse
72 0803 0.84 0.51 347.90 1.3 "Loma Prieta" 1989 Coll." 6.93 Oblique 8.48 9.31
strike
73 0821 2.42 1.87 352.05 1.8 "Erzican Turkey" 1992 "Erzincan" 6.69 slip 0 4.38
strike
74 0821 1.08 0.83 352.05 1.8 "Erzican Turkey" 1992 "Erzincan" 6.69 slip 0 4.38
75 0825 2.29 1.39 567.78 6 "Cape Mendocino" 1992 "Cape Mendocino" 7.01 Reverse 0 6.96
76 0825 1.12 0.68 567.78 6 "Cape Mendocino" 1992 "Cape Mendocino" 7.01 Reverse 0 6.96
19.9
77 0827 0.95 0.17 457.06 0.3 "Cape Mendocino" 1992 "Fortuna - Fortuna Blvd" 7.01 Reverse 15.97 5
78 0828 2.36 1.93 422.17 3.8 "Cape Mendocino" 1992 "Petrolia" 7.01 Reverse 0 8.18
79 0828 1.05 0.86 422.17 3.8 "Cape Mendocino" 1992 "Petrolia" 7.01 Reverse 0 8.18
strike 68.6
80 0860 1.16 0.11 328.09 0.3 "Landers" 1992 "Hemet Fire Station" 7.28 slip 68.66 6
strike 26.9
81 0880 1.01 0.09 355.42 0.4 "Landers" 1992 "Mission Creek Fault" 7.28 slip 26.96 6
"Morongo Valley Fire strike 17.3
82 0881 0.94 0.20 396.41 1.2 "Landers" 1992 Station" 7.28 slip 17.36 6
strike 41.4
83 0897 1.04 0.03 635.01 0.1 "Landers" 1992 "Twentynine Palms" 7.28 slip 41.43 3
strike 23.6
84 0900 2.18 0.92 353.63 0.9 "Landers" 1992 "Yermo Fire Station" 7.28 slip 23.62 2
strike 23.6
85 0900 0.88 0.37 353.63 0.9 "Landers" 1992 "Yermo Fire Station" 7.28 slip 23.62 2
"Beverly Hills - 12520 18.3
86 0952 0.88 0.26 545.66 3 "Northridge-01" 1994 Mulhol" 6.69 Reverse 12.39 6
"Beverly Hills - 14145 17.1
87 0953 1.18 1.15 355.81 4.5 "Northridge-01" 1994 Mulhol" 6.69 Reverse 9.44 5
53.4
88 0966 1.00 0.08 324.79 0.1 "Northridge-01" 1994 "Covina - W Badillo" 6.69 Reverse 53.21 5
"Downey - Co Maint 46.7
89 0968 0.97 0.15 271.90 0.6 "Northridge-01" 1994 Bldg" 6.69 Reverse 43.2 4
193
53.9
90 0975 0.91 0.09 362.31 0.1 "Northridge-01" 1994 "Glendora - N Oakbank" 6.69 Reverse 53.71 4
"Jensen Filter Plant
91 0982 1.74 2.48 373.07 5.3 "Northridge-01" 1994 Administrative Building" 6.69 Reverse 0 5.43
"Jensen Filter Plant
92 0982 0.93 1.32 373.07 5.3 "Northridge-01" 1994 Administrative Building" 6.69 Reverse 0 5.43
"Jensen Filter Plant
93 0983 1.93 1.93 525.79 6.5 "Northridge-01" 1994 Generator Building" 6.69 Reverse 0 5.43
"Jensen Filter Plant
94 0983 1.07 1.07 525.79 6.5 "Northridge-01" 1994 Generator Building" 6.69 Reverse 0 5.43
41.1
95 0984 1.05 0.14 301.00 0.4 "Northridge-01" 1994 "LA - 116th St School" 6.69 Reverse 36.39 7
36.6
96 0990 0.98 0.15 365.22 1.1 "Northridge-01" 1994 "LA - City Terrace" 6.69 Reverse 35.03 2
26.7
97 0998 1.00 0.18 315.06 1.4 "Northridge-01" 1994 "LA - N Westmoreland" 6.69 Reverse 23.4 3
33.9
98 1001 0.98 0.19 285.28 0.7 "Northridge-01" 1994 "LA - S Grand Ave" 6.69 Reverse 29.52 9
"LA - Sepulveda VA
99 1004 1.66 1.42 380.06 7 "Northridge-01" 1994 Hospital" 6.69 Reverse 0 8.44
"LA - Sepulveda VA
100 1004 0.92 0.78 380.06 7 "Northridge-01" 1994 Hospital" 6.69 Reverse 0 8.44
22.4
101 1006 1.09 0.25 398.42 1.6 "Northridge-01" 1994 "LA - UCLA Grounds" 6.69 Reverse 13.8 9
102 1013 2.33 1.46 628.99 1.8 "Northridge-01" 1994 "LA Dam" 6.69 Reverse 0 5.92
103 1013 1.13 0.71 628.99 1.8 "Northridge-01" 1994 "LA Dam" 6.69 Reverse 0 5.92
"Mojave - Oak Creek
104 1037 0.96 0.03 422.73 0 "Northridge-01" 1994 Canyon" 6.69 Reverse 75.64 75.8
105 1044 1.71 1.71 269.14 5.7 "Northridge-01" 1994 "Newhall - Fire Sta" 6.69 Reverse 3.16 5.92
106 1044 0.91 0.91 269.14 5.7 "Northridge-01" 1994 "Newhall - Fire Sta" 6.69 Reverse 3.16 5.92
107 1052 0.97 0.50 508.08 1.8 "Northridge-01" 1994 "Pacoima Kagel Canyon" 6.69 Reverse 5.26 7.26
108 1054 2.12 2.47 325.67 3.1 "Northridge-01" 1994 "Pardee - SCE" 6.69 Reverse 5.54 7.46
109 1054 1.17 1.37 325.67 3.1 "Northridge-01" 1994 "Pardee - SCE" 6.69 Reverse 5.54 7.46
"Rancho Palos Verdes - 52.1
110 1061 1.14 0.07 580.03 0.1 "Northridge-01" 1994 Hawth" 6.69 Reverse 48.02 8
194
111 1063 1.86 2.72 282.25 7.5 "Northridge-01" 1994 "Rinaldi Receiving Sta" 6.69 Reverse 0 6.5
112 1063 0.91 1.33 282.25 7.5 "Northridge-01" 1994 "Rinaldi Receiving Sta" 6.69 Reverse 0 6.5
"Simi Valley - Katherine 13.4
113 1080 2.32 1.66 557.42 4.1 "Northridge-01" 1994 Rd" 6.69 Reverse 0 2
"Simi Valley - Katherine 13.4
114 1080 1.06 0.76 557.42 4.1 "Northridge-01" 1994 Rd" 6.69 Reverse 0 2
115 1084 1.62 2.24 251.24 6 "Northridge-01" 1994 "Sylmar - Converter Sta" 6.69 Reverse 0 5.35
116 1084 1.01 1.40 251.24 6 "Northridge-01" 1994 "Sylmar - Converter Sta" 6.69 Reverse 0 5.35
"Sylmar - Olive View
117 1086 1.75 1.14 440.54 5 "Northridge-01" 1994 Med FF" 6.69 Reverse 1.74 5.3
"Sylmar - Olive View
118 1086 0.97 1.12 440.54 5 "Northridge-01" 1994 Med FF" 6.69 Reverse 1.74 5.3
"Wrightwood - Nielson 81.6
119 1097 0.98 0.03 506.00 0 "Northridge-01" 1994 Ranch" 6.69 Reverse 81.54 9
strike 11.3
120 1101 1.87 1.58 256.00 2 "Kobe Japan" 1995 "Amagasaki" 6.9 slip 11.34 4
strike 11.3
121 1101 0.97 0.82 256.00 2 "Kobe Japan" 1995 "Amagasaki" 6.9 slip 11.34 4
strike
122 1106 1.69 2.34 312.00 8.4 "Kobe Japan" 1995 "KJMA" 6.9 slip 0.94 0.96
strike
123 1106 1.16 1.61 312.00 8.4 "Kobe Japan" 1995 "KJMA" 6.9 slip 0.94 0.96
strike
124 1107 0.97 0.33 312.00 1.7 "Kobe Japan" 1995 "Kakogawa" 6.9 slip 22.5 22.5
strike 70.2
125 1109 0.91 0.03 609.00 0.1 "Kobe Japan" 1995 "MZH" 6.9 slip 69.04 6
strike
126 1111 2.31 0.66 609.00 3.4 "Kobe Japan" 1995 "Nishi-Akashi" 6.9 slip 7.08 7.08
strike
127 1111 1.06 0.30 609.00 3.4 "Kobe Japan" 1995 "Nishi-Akashi" 6.9 slip 7.08 7.08
strike
128 1114 2.31 2.15 198.00 1.8 "Kobe Japan" 1995 "Port Island (0 m)" 6.9 slip 3.31 3.31
strike
129 1114 1.13 1.05 198.00 1.8 "Kobe Japan" 1995 "Port Island (0 m)" 6.9 slip 3.31 3.31
strike 28.0
130 1115 1.02 0.18 256.00 0.6 "Kobe Japan" 1995 "Sakai" 6.9 slip 28.08 8
195
strike 19.1
131 1116 1.02 0.27 256.00 0.8 "Kobe Japan" 1995 "Shin-Osaka" 6.9 slip 19.14 5
strike
132 1119 2.27 1.86 312.00 3.9 "Kobe Japan" 1995 "Takarazuka" 6.9 slip 0 0.27
strike
133 1119 1.11 0.91 312.00 3.9 "Kobe Japan" 1995 "Takarazuka" 6.9 slip 0 0.27
strike
134 1120 1.62 2.09 256.00 8.7 "Kobe Japan" 1995 "Takatori" 6.9 slip 1.46 1.47
strike
135 1120 1.02 1.31 256.00 8.7 "Kobe Japan" 1995 "Takatori" 6.9 slip 1.46 1.47
strike 27.7
136 1121 0.91 0.37 256.00 1.1 "Kobe Japan" 1995 "Yae" 6.9 slip 27.77 7
strike 65.5
137 1154 1.00 0.12 612.78 0.1 "Kocaeli Turkey" 1999 "Bursa Sivil" 7.51 slip 65.53 3
strike 15.3
138 1158 2.01 0.98 281.86 1.3 "Kocaeli Turkey" 1999 "Duzce" 7.51 slip 13.6 7
strike 15.3
139 1158 0.92 0.45 281.86 1.3 "Kocaeli Turkey" 1999 "Duzce" 7.51 slip 13.6 7
strike 31.7
140 1162 1.06 0.14 347.62 0.3 "Kocaeli Turkey" 1999 "Goynuk" 7.51 slip 31.74 4
strike 30.7
141 1166 0.94 0.21 476.62 0.4 "Kocaeli Turkey" 1999 "Iznik" 7.51 slip 30.73 3
strike
142 1176 2.33 0.90 297.00 1.3 "Kocaeli Turkey" 1999 "Yarimca" 7.51 slip 1.38 4.83
strike
143 1176 1.14 0.44 297.00 1.3 "Kocaeli Turkey" 1999 "Yarimca" 7.51 slip 1.38 4.83
Reverse
144 1197 1.49 1.51 542.61 5.9 "Chi-Chi Taiwan" 1999 "CHY028" 7.62 Oblique 3.12 3.12
Reverse
145 1197 1.02 1.04 542.61 5.9 "Chi-Chi Taiwan" 1999 "CHY028" 7.62 Oblique 3.12 3.12
Reverse
146 1231 1.11 2.34 496.21 9.3 "Chi-Chi Taiwan" 1999 "CHY080" 7.62 Oblique 0.11 2.69
Reverse 28.4
147 1234 0.92 0.21 665.20 1 "Chi-Chi Taiwan" 1999 "CHY086" 7.62 Oblique 27.57 2
Reverse
148 1244 2.35 1.73 258.89 3 "Chi-Chi Taiwan" 1999 "CHY101" 7.62 Oblique 9.94 9.94
196
Reverse
149 1244 1.04 0.77 258.89 3 "Chi-Chi Taiwan" 1999 "CHY101" 7.62 Oblique 9.94 9.94
Reverse 47.7
150 1289 1.07 0.26 484.97 0.3 "Chi-Chi Taiwan" 1999 "HWA041" 7.62 Oblique 43.37 6
Reverse 16.7
151 1486 1.10 0.18 465.55 0.4 "Chi-Chi Taiwan" 1999 "TCU046" 7.62 Oblique 16.74 4
Reverse
152 1492 2.22 2.27 579.10 2.9 "Chi-Chi Taiwan" 1999 "TCU052" 7.62 Oblique 0 0.66
Reverse
153 1492 0.96 0.98 579.10 2.9 "Chi-Chi Taiwan" 1999 "TCU052" 7.62 Oblique 0 0.66
Reverse
154 1503 1.90 2.22 305.85 7.7 "Chi-Chi Taiwan" 1999 "TCU065" 7.62 Oblique 0.57 0.57
Reverse
155 1503 0.93 1.09 305.85 7.7 "Chi-Chi Taiwan" 1999 "TCU065" 7.62 Oblique 0.57 0.57
Reverse
156 1505 1.51 1.06 487.34 3.3 "Chi-Chi Taiwan" 1999 "TCU068" 7.62 Oblique 0 0.32
Reverse
157 1505 1.04 0.73 487.34 3.3 "Chi-Chi Taiwan" 1999 "TCU068" 7.62 Oblique 0 0.32
Reverse
158 1507 2.05 1.43 624.85 9.5 "Chi-Chi Taiwan" 1999 "TCU071" 7.62 Oblique 0 5.8
Reverse
159 1507 1.00 0.70 624.85 9.5 "Chi-Chi Taiwan" 1999 "TCU071" 7.62 Oblique 0 5.8
Reverse 13.4
160 1509 1.85 2.11 549.43 6.4 "Chi-Chi Taiwan" 1999 "TCU074" 7.62 Oblique 0 6
Reverse 13.4
161 1509 0.90 1.03 549.43 6.4 "Chi-Chi Taiwan" 1999 "TCU074" 7.62 Oblique 0 6
Reverse
162 1510 1.99 0.69 573.02 3 "Chi-Chi Taiwan" 1999 "TCU075" 7.62 Oblique 0.89 0.89
Reverse
163 1510 1.03 0.36 573.02 3 "Chi-Chi Taiwan" 1999 "TCU075" 7.62 Oblique 0.89 0.89
Reverse 10.9
164 1513 1.39 0.88 363.99 7.7 "Chi-Chi Taiwan" 1999 "TCU079" 7.62 Oblique 0 7
Reverse 10.9
165 1513 0.93 0.59 363.99 7.7 "Chi-Chi Taiwan" 1999 "TCU079" 7.62 Oblique 0 7
Reverse 11.4
166 1517 1.06 1.99 665.20 20.3 "Chi-Chi Taiwan" 1999 "TCU084" 7.62 Oblique 0 8
197
Reverse
167 1549 2.42 1.37 511.18 9.3 "Chi-Chi Taiwan" 1999 "TCU129" 7.62 Oblique 1.83 1.83
Reverse
168 1549 0.95 0.54 511.18 9.3 "Chi-Chi Taiwan" 1999 "TCU129" 7.62 Oblique 1.83 1.83
Reverse
169 1551 1.04 0.45 652.85 1.7 "Chi-Chi Taiwan" 1999 "TCU138" 7.62 Oblique 9.78 9.78
strike 12.0
170 1602 2.23 2.16 293.57 3.7 "Duzce Turkey" 1999 "Bolu" 7.14 slip 12.02 4
strike 12.0
171 1602 1.09 1.05 293.57 3.7 "Duzce Turkey" 1999 "Bolu" 7.14 slip 12.02 4
strike
172 1605 2.36 1.51 281.86 2.9 "Duzce Turkey" 1999 "Duzce" 7.14 slip 0 6.58
strike
173 1605 1.15 0.74 281.86 2.9 "Duzce Turkey" 1999 "Duzce" 7.14 slip 0 6.58
strike 45.1
174 1620 1.12 0.02 411.91 0 "Duzce Turkey" 1999 "Sakarya" 7.14 slip 45.16 6
strike 34.6
175 1626 1.07 0.05 649.67 0.2 "Sitka Alaska" 1972 "Sitka Observatory" 7.68 slip 34.61 1
strike 50.8
176 1627 1.07 0.03 432.58 0.1 "Caldiran Turkey" 1976 "Maku" 7.21 slip 50.78 2
26.4
177 1628 0.97 0.27 306.37 0.9 "St Elias Alaska" 1979 "Icy Bay" 7.54 Reverse 26.46 6
strike 49.9
178 1636 1.08 0.13 302.64 0.4 "Manjil Iran" 1990 "Qazvin" 7.37 slip 49.97 7
"Banning - Twin Pines strike 83.4
179 1767 0.97 0.02 667.42 0 "Hector Mine" 1999 Road" 7.13 slip 83.43 3
strike 74.9
180 1782 1.03 0.08 436.14 0.1 "Hector Mine" 1999 "Forest Falls Post Office" 7.13 slip 74.92 2
strike 31.0
181 1794 0.92 0.28 379.32 0.6 "Hector Mine" 1999 "Joshua Tree" 7.13 slip 31.06 6
"Nenana Mountain "TAPS Pump Station strike 104.7 104.
182 2093 1.08 0.02 382.50 0 Alaska" 2002 #09" 6.7 slip 3 73
strike
183 2111 0.88 0.09 341.56 0.1 "Denali Alaska" 2002 "R109 (temp)" 7.9 slip 42.99 43
"TAPS Pump Station strike
184 2114 2.40 1.79 329.40 1.9 "Denali Alaska" 2002 #10" 7.9 slip 0.18 2.74
198
"TAPS Pump Station strike
185 2114 1.17 0.87 329.40 1.9 "Denali Alaska" 2002 #10" 7.9 slip 0.18 2.74
"Taiwan 95.9
186 3583 1.22 0.07 309.41 0 SMART1(25)" 1983 "SMART1 I08" 6.5 Reverse 95.5 8
"Taiwan
187 3594 1.04 0.06 300.22 0 SMART1(25)" 1983 "SMART1 M11" 6.5 Reverse 96.52 97
12.2
188 3744 1.06 0.40 566.42 0.6 "Cape Mendocino" 1992 "Bunker Hill FAA" 7.01 Reverse 8.49 4
"Centerville Beach Naval 18.3
189 3746 2.23 0.97 459.04 1.6 "Cape Mendocino" 1992 Fac" 7.01 Reverse 16.44 1
"Centerville Beach Naval 18.3
190 3746 1.02 0.44 459.04 1.6 "Cape Mendocino" 1992 Fac" 7.01 Reverse 16.44 1
19.3
191 3748 2.48 1.63 387.95 1.7 "Cape Mendocino" 1992 "Ferndale Fire Station" 7.01 Reverse 16.64 2
19.3
192 3748 1.21 0.80 387.95 1.7 "Cape Mendocino" 1992 "Ferndale Fire Station" 7.01 Reverse 16.64 2
20.4
193 3749 2.06 0.68 355.18 1.3 "Cape Mendocino" 1992 "Fortuna Fire Station" 7.01 Reverse 16.54 1
20.4
194 3749 0.98 0.32 355.18 1.3 "Cape Mendocino" 1992 "Fortuna Fire Station" 7.01 Reverse 16.54 1
25.9
195 3750 2.08 0.51 515.65 0.9 "Cape Mendocino" 1992 "Loleta Fire Station" 7.01 Reverse 23.46 1
25.9
196 3750 0.83 0.20 515.65 0.9 "Cape Mendocino" 1992 "Loleta Fire Station" 7.01 Reverse 23.46 1
"Thousand Palms Post strike 36.9
197 3758 1.01 0.20 333.89 0.5 "Landers" 1992 Office" 7.28 slip 36.93 3
strike 82.4
198 3882 1.22 0.02 571.63 0 "Tottori Japan" 2000 "HRS016" 6.61 slip 82.42 2
strike 88.7
199 3899 1.00 0.01 617.44 0 "Tottori Japan" 2000 "HYGH02" 6.61 slip 88.75 5
strike 28.8
200 3908 1.07 0.13 293.37 0.8 "Tottori Japan" 2000 "OKY005" 6.61 slip 28.81 2
strike 66.2
201 3915 1.23 0.08 296.96 0.1 "Tottori Japan" 2000 "OKY012" 6.61 slip 66.24 5
strike 45.7
202 3937 1.09 0.11 182.30 0.2 "Tottori Japan" 2000 "SMN005" 6.61 slip 45.73 3
199
strike 77.8
203 3945 0.86 0.02 262.19 0 "Tottori Japan" 2000 "SMN017" 6.61 slip 77.85 5
strike 85.3
204 3946 0.99 0.05 271.29 0.1 "Tottori Japan" 2000 "SMN018" 6.61 slip 85.31 1
strike
205 3968 1.84 2.58 310.21 11.8 "Tottori Japan" 2000 "TTRH02" 6.61 slip 0.83 0.97
strike
206 3968 1.02 1.43 310.21 11.8 "Tottori Japan" 2000 "TTRH02" 6.61 slip 0.83 0.97
"Coalinga - Fire Station 70.2
207 3981 0.86 0.05 333.61 0 "San Simeon CA" 2003 39" 6.52 Reverse 69.51 3
"Greenfield - Police
208 3987 0.87 0.03 280.64 0 "San Simeon CA" 2003 Station" 6.52 Reverse 69.08 69.8
"San Luis Obispo - 48.1
209 3994 1.05 0.10 365.15 0.2 "San Simeon CA" 2003 Lopez Lake Grounds" 6.52 Reverse 48.07 1
"Templeton - 1-story
210 4031 2.28 0.76 410.66 1.9 "San Simeon CA" 2003 Hospital" 6.52 Reverse 5.07 6.22
"Templeton - 1-story
211 4031 0.96 0.32 410.66 1.9 "San Simeon CA" 2003 Hospital" 6.52 Reverse 5.07 6.22
strike
212 4040 2.28 1.74 487.40 8 "Bam Iran" 2003 "Bam" 6.6 slip 0.05 1.7
strike
213 4040 0.99 0.75 487.40 8 "Bam Iran" 2003 "Bam" 6.6 slip 0.05 1.7
"Mohammad Abad-e- strike 46.2
214 4054 0.83 0.04 574.88 0.2 "Bam Iran" 2003 Madkoon" 6.6 slip 46.2 2
84.2
215 4198 0.98 0.02 220.65 0 "Niigata Japan" 2004 "NIG008" 6.63 Reverse 83.83 8
12.8
216 4207 0.98 0.33 274.17 3.4 "Niigata Japan" 2004 "NIG017" 6.63 Reverse 4.22 1
25.8
217 4208 0.91 0.14 198.26 0.8 "Niigata Japan" 2004 "NIG018" 6.63 Reverse 21.55 4
18.0
218 4212 1.10 0.13 193.20 0.8 "Niigata Japan" 2004 "NIG022" 6.63 Reverse 17.57 3
219 4218 0.96 0.32 430.71 5.2 "Niigata Japan" 2004 "NIG028" 6.63 Reverse 0.46 9.79
220 4219 2.25 1.72 480.40 8.8 "Niigata Japan" 2004 "NIGH01" 6.63 Reverse 0.49 9.46
221 4219 1.10 0.84 480.40 8.8 "Niigata Japan" 2004 "NIGH01" 6.63 Reverse 0.49 9.46
200
71.5
222 4222 1.05 0.04 244.84 0.2 "Niigata Japan" 2004 "NIGH05" 6.63 Reverse 70.59 2
223 4228 2.42 0.96 375.00 2.2 "Niigata Japan" 2004 "NIGH11" 6.63 Reverse 6.27 8.93
224 4228 1.11 0.44 375.00 2.2 "Niigata Japan" 2004 "NIGH11" 6.63 Reverse 6.27 8.93
"Montenegro
225 4451 1.97 1.71 462.23 3 Yugoslavia" 1979 "Bar-Skupstina Opstine" 7.1 Reverse 0 6.98
"Montenegro
226 4451 1.23 1.07 462.23 3 Yugoslavia" 1979 "Bar-Skupstina Opstine" 7.1 Reverse 0 6.98
"Montenegro
227 4456 0.93 0.42 543.26 4.6 Yugoslavia" 1979 "Petrovac - Hotel Olivia" 7.1 Reverse 0 8.01
"Montenegro
228 4458 1.95 1.06 318.74 1.8 Yugoslavia" 1979 "Ulcinj - Hotel Olimpic" 7.1 Reverse 3.97 5.76
"Montenegro
229 4458 1.05 0.57 318.74 1.8 Yugoslavia" 1979 "Ulcinj - Hotel Olimpic" 7.1 Reverse 3.97 5.76
"Joetsu Uragawaraku 22.7
230 4842 0.96 0.17 655.45 1.4 "Chuetsu-oki Japan" 2007 Kamabucchi" 6.8 Reverse 18.6 4
"Tokamachi 28.7
231 4844 0.93 0.18 640.14 0.3 "Chuetsu-oki Japan" 2007 Matsunoyama" 6.8 Reverse 23.01 5
"Kubikiku Hyakken 22.1
232 4849 0.96 0.36 342.74 0.8 "Chuetsu-oki Japan" 2007 Joetsu City" 6.8 Reverse 20.71 8
"Kashiwazaki City 11.0
233 4856 2.17 1.80 294.38 3.9 "Chuetsu-oki Japan" 2007 Center" 6.8 Reverse 0 9
"Kashiwazaki City 11.0
234 4856 0.93 0.78 294.38 3.9 "Chuetsu-oki Japan" 2007 Center" 6.8 Reverse 0 9
"Mitsuke Kazuiti Arita 20.3
235 4859 0.95 0.37 274.23 0.8 "Chuetsu-oki Japan" 2007 Town" 6.8 Reverse 11.35 3
16.2
236 4863 2.00 1.35 514.30 2.2 "Chuetsu-oki Japan" 2007 "Nagaoka" 6.8 Reverse 3.97 7
16.2
237 4863 1.17 0.79 514.30 2.2 "Chuetsu-oki Japan" 2007 "Nagaoka" 6.8 Reverse 3.97 7
"Sawa Mizuguti
238 4872 1.04 0.27 640.14 0.3 "Chuetsu-oki Japan" 2007 Tokamachi" 6.8 Reverse 21.17 27.3
239 4874 2.42 1.28 561.59 5.3 "Chuetsu-oki Japan" 2007 "Oguni Nagaoka" 6.8 Reverse 10.31 20
240 4874 1.18 0.62 561.59 5.3 "Chuetsu-oki Japan" 2007 "Oguni Nagaoka" 6.8 Reverse 10.31 20
241 4875 1.08 0.89 282.57 6.4 "Chuetsu-oki Japan" 2007 "Kariwa" 6.8 Reverse 0 12
201
"Kashiwazaki 12.6
242 4876 2.11 1.98 655.45 8.6 "Chuetsu-oki Japan" 2007 Nishiyamacho Ikeura" 6.8 Reverse 0 3
"Kashiwazaki 12.6
243 4876 1.03 0.96 655.45 8.6 "Chuetsu-oki Japan" 2007 Nishiyamacho Ikeura" 6.8 Reverse 0 3
"Yan Sakuramachi City 18.9
244 4879 1.09 0.57 265.82 0.7 "Chuetsu-oki Japan" 2007 watershed" 6.8 Reverse 12.98 7
"Tamati Yone 11.4
245 4886 2.22 1.19 338.32 4.9 "Chuetsu-oki Japan" 2007 Izumozaki" 6.8 Reverse 0 8
"Tamati Yone 11.4
246 4886 1.08 0.58 338.32 4.9 "Chuetsu-oki Japan" 2007 Izumozaki" 6.8 Reverse 0 8
"Kashiwazaki NPP Unit 10.9
247 4894 1.36 2.15 329.00 16.5 "Chuetsu-oki Japan" 2007 1: ground surface" 6.8 Reverse 0 7
"Kashiwazaki NPP Unit 10.9
248 4894 0.97 1.53 329.00 16.5 "Chuetsu-oki Japan" 2007 1: ground surface" 6.8 Reverse 0 7
"Kashiwazaki NPP Unit 10.9
249 4895 1.33 1.51 265.50 13.3 "Chuetsu-oki Japan" 2007 5: ground surface" 6.8 Reverse 0 7
"Kashiwazaki NPP Unit 10.9
250 4895 1.03 1.05 265.50 13.3 "Chuetsu-oki Japan" 2007 5: ground surface" 6.8 Reverse 0 7
"Kashiwazaki NPP
Service Hall Array 2.4 m 10.9
251 4896 0.93 0.91 201.00 5.1 "Chuetsu-oki Japan" 2007 depth" 6.8 Reverse 0 7
55.3
252 4997 1.00 0.09 305.54 0.1 "Chuetsu-oki Japan" 2007 "FKS028" 6.8 Reverse 52.63 8
95.0
253 5003 0.80 0.01 245.88 0 "Chuetsu-oki Japan" 2007 "FKSH04" 6.8 Reverse 93.48 5
87.9
254 5064 1.03 0.03 342.36 0 "Chuetsu-oki Japan" 2007 "GNM005" 6.8 Reverse 86.23 4
83.3
255 5254 0.96 0.02 220.65 0 "Chuetsu-oki Japan" 2007 "NIG008" 6.8 Reverse 81.51 1
67.7
256 5258 1.00 0.07 229.95 0.3 "Chuetsu-oki Japan" 2007 "NIG012" 6.8 Reverse 65.54 7
10.7
257 5264 1.77 1.66 198.26 5 "Chuetsu-oki Japan" 2007 "NIG018" 6.8 Reverse 0 8
10.7
258 5264 1.11 1.04 198.26 5 "Chuetsu-oki Japan" 2007 "NIG018" 6.8 Reverse 0 8
202
112.7 113.
259 5461 0.89 0.02 279.36 0 "Iwate Japan" 2008 "AKT006" 6.9 Reverse 8 45
58.6
260 5467 0.98 0.02 449.45 0.1 "Iwate Japan" 2008 "AKT012" 6.9 Reverse 57.37 7
48.3
261 5471 1.08 0.09 158.16 0.3 "Iwate Japan" 2008 "AKT016" 6.9 Reverse 46.77 6
96.1
262 5490 1.14 0.01 232.58 0.1 "Iwate Japan" 2008 "AKTH14" 6.9 Reverse 95.32 1
49.9
263 5648 1.12 0.04 534.71 0.1 "Iwate Japan" 2008 "IWTH16" 6.9 Reverse 48.43 7
264 5656 2.34 0.78 486.41 3.5 "Iwate Japan" 2008 "IWTH24" 6.9 Reverse 3.1 5.18
265 5656 1.14 0.38 486.41 3.5 "Iwate Japan" 2008 "IWTH24" 6.9 Reverse 3.1 5.18
266 5657 1.85 1.40 506.44 26.1 "Iwate Japan" 2008 "IWTH25" 6.9 Reverse 0 4.8
267 5657 1.02 0.78 506.44 26.1 "Iwate Japan" 2008 "IWTH25" 6.9 Reverse 0 4.8
268 5658 2.37 1.06 371.06 14.1 "Iwate Japan" 2008 "IWTH26" 6.9 Reverse 5.97 6.02
269 5658 1.15 0.52 371.06 14.1 "Iwate Japan" 2008 "IWTH26" 6.9 Reverse 5.97 6.02
20.1
270 5663 2.38 0.96 479.37 9.4 "Iwate Japan" 2008 "MYG004" 6.9 Reverse 20.17 8
20.1
271 5663 1.03 0.41 479.37 9.4 "Iwate Japan" 2008 "MYG004" 6.9 Reverse 20.17 8
13.4
272 5664 2.38 1.07 361.24 4.2 "Iwate Japan" 2008 "MYG005" 6.9 Reverse 10.71 7
13.4
273 5664 1.16 0.52 361.24 4.2 "Iwate Japan" 2008 "MYG005" 6.9 Reverse 10.71 7
48.5
274 5768 0.99 0.03 291.48 0.1 "Iwate Japan" 2008 "YMTH09" 6.9 Reverse 47.01 9
29.3
275 5774 0.94 0.19 276.30 1 "Iwate Japan" 2008 "Nakashinden Town" 6.9 Reverse 29.37 8
20.7
276 5780 1.91 0.81 345.55 1.8 "Iwate Japan" 2008 "Iwadeyama" 6.9 Reverse 20.77 8
20.7
277 5780 0.91 0.39 345.55 1.8 "Iwate Japan" 2008 "Iwadeyama" 6.9 Reverse 20.77 8
"Misato Akita City - 41.7
278 5799 1.04 0.08 552.38 0.4 "Iwate Japan" 2008 Tsuchizaki" 6.9 Reverse 39.86 2
203
12.8
279 5818 2.35 1.24 512.26 7.3 "Iwate Japan" 2008 "Kurihara City" 6.9 Reverse 12.83 5
12.8
280 5818 1.05 0.55 512.26 7.3 "Iwate Japan" 2008 "Kurihara City" 6.9 Reverse 12.83 5
"El Mayor-Cucapah "CERRO PRIETO strike 10.9
281 5825 2.34 0.91 242.05 3.3 Mexico" 2010 GEOTHERMAL" 7.2 slip 8.88 2
"El Mayor-Cucapah "CERRO PRIETO strike 10.9
282 5825 0.94 0.37 242.05 3.3 Mexico" 2010 GEOTHERMAL" 7.2 slip 8.88 2
"El Mayor-Cucapah "MICHOACAN DE strike 15.9
283 5827 2.35 1.38 242.05 6.1 Mexico" 2010 OCAMPO" 7.2 slip 13.21 1
"El Mayor-Cucapah "MICHOACAN DE strike 15.9
284 5827 1.15 0.67 242.05 6.1 Mexico" 2010 OCAMPO" 7.2 slip 13.21 1
"El Mayor-Cucapah "El Centro - Imperial and strike 20.0
285 5837 2.27 1.22 229.25 3.7 Mexico" 2010 Ross" 7.2 slip 19.39 8
"El Mayor-Cucapah "El Centro - Imperial and strike 20.0
286 5837 0.92 0.49 229.25 3.7 Mexico" 2010 Ross" 7.2 slip 19.39 8
"El Mayor-Cucapah strike
287 5839 1.01 0.02 388.01 0 Mexico" 2010 "El Cajon - Marshall" 7.2 slip 115 115
"El Mayor-Cucapah strike
288 5864 1.01 0.08 384.66 0 Mexico" 2010 "Frink" 7.2 slip 81.63 81.8
"El Mayor-Cucapah strike
289 5970 0.82 0.01 619.00 0 Mexico" 2010 "Borrego Springs" 7.2 slip 91.9 91.9
"El Mayor-Cucapah strike 41.4
290 5972 0.91 0.11 208.71 0.8 Mexico" 2010 "Brawley Airport" 7.2 slip 41.15 8
"El Mayor-Cucapah strike 20.4
291 5975 1.87 0.60 231.23 2.4 Mexico" 2010 "Calexico Fire Station" 7.2 slip 19.12 6
"El Mayor-Cucapah strike 20.4
292 5975 0.89 0.29 231.23 2.4 Mexico" 2010 "Calexico Fire Station" 7.2 slip 19.12 6
"El Mayor-Cucapah "El Centro Differential strike 23.4
293 5985 2.19 1.22 202.26 4.3 Mexico" 2010 Array" 7.2 slip 22.83 2
"El Mayor-Cucapah "El Centro Differential strike 23.4
294 5985 0.81 0.45 202.26 4.3 Mexico" 2010 Array" 7.2 slip 22.83 2
"El Mayor-Cucapah strike 20.0
295 5991 1.76 1.01 202.85 3.6 Mexico" 2010 "El Centro Array #10" 7.2 slip 19.36 5
"El Mayor-Cucapah strike 20.0
296 5991 1.10 0.63 202.85 3.6 Mexico" 2010 "El Centro Array #10" 7.2 slip 19.36 5
204
"El Mayor-Cucapah strike 16.2
297 5992 2.50 1.51 196.25 5.5 Mexico" 2010 "El Centro Array #11" 7.2 slip 15.36 1
"El Mayor-Cucapah strike 16.2
298 5992 1.08 0.65 196.25 5.5 Mexico" 2010 "El Centro Array #11" 7.2 slip 15.36 1
111.3 111.
299 6515 0.95 0.02 279.58 0 "Niigata Japan" 2004 "FKS016" 6.63 Reverse 3 4
109.1 109.
300 6783 1.01 0.02 265.60 0 "Niigata Japan" 2004 "TCG008" 6.63 Reverse 4 21
"Darfield New strike 14.4
301 6886 1.00 0.16 280.26 0.9 Zealand" 2010 "Canterbury Aero Club" 7 slip 14.48 8
"Darfield New strike 11.8
302 6893 2.14 0.86 344.02 2.8 Zealand" 2010 "DFHS" 7 slip 11.86 6
"Darfield New strike 11.8
303 6893 1.11 0.44 344.02 2.8 Zealand" 2010 "DFHS" 7 slip 11.86 6
"Darfield New strike
304 6906 1.79 1.82 344.02 4.7 Zealand" 2010 "GDLC" 7 slip 1.22 1.22
"Darfield New strike
305 6906 1.12 1.14 344.02 4.7 Zealand" 2010 "GDLC" 7 slip 1.22 1.22
"Darfield New strike
306 6911 2.04 1.42 326.01 3.2 Zealand" 2010 "HORC" 7 slip 7.29 7.29
"Darfield New strike
307 6911 1.13 0.79 326.01 3.2 Zealand" 2010 "HORC" 7 slip 7.29 7.29
"Darfield New strike
308 6927 2.26 1.28 263.20 2.7 Zealand" 2010 "LINC" 7 slip 5.07 7.11
"Darfield New strike
309 6927 1.11 0.62 263.20 2.7 Zealand" 2010 "LINC" 7 slip 5.07 7.11
"Darfield New strike 25.6
310 6928 0.98 0.17 649.67 0.7 Zealand" 2010 "LPCC" 7 slip 25.21 7
"Darfield New strike 35.2
311 6933 1.09 0.05 342.70 0.1 Zealand" 2010 "MAYC" 7 slip 33.54 3
"Darfield New "Pages Road Pumping strike 24.5
312 6953 2.16 0.64 206.00 1.3 Zealand" 2010 Station" 7 slip 24.55 5
"Darfield New "Pages Road Pumping strike 24.5
313 6953 1.03 0.30 206.00 1.3 Zealand" 2010 Station" 7 slip 24.55 5
"Darfield New strike
314 6962 2.23 0.85 295.74 1.6 Zealand" 2010 "ROLC" 7 slip 0 1.54
205
"Darfield New strike
315 6962 1.09 0.42 295.74 1.6 Zealand" 2010 "ROLC" 7 slip 0 1.54
"Darfield New strike 24.3
316 6965 0.95 0.12 263.20 0.7 Zealand" 2010 "SBRC" 7 slip 21.31 4
"El Mayor-Cucapah strike 11.2
317 8161 2.49 1.67 196.88 3.2 Mexico" 2010 "El Centro Array #12" 7.2 slip 9.98 6
"El Mayor-Cucapah strike 11.2
318 8161 1.22 0.81 196.88 3.2 Mexico" 2010 "El Centro Array #12" 7.2 slip 9.98 6
"El Mayor-Cucapah "SANTA ISABEL strike 57.4
319 8163 1.02 0.02 483.02 0 Mexico" 2010 VIEJO" 7.2 slip 55.19 9
strike
320 8166 1.01 0.19 425.00 - "Duzce Turkey" 1999 "IRIGM 498" 7.14 slip 3.58 3.58
206
APPENDIX B. FRAGILITY CURVES IN TERMS OF PGA
Chapter 6 presented the approach and methodology for the generation of fragility
curves. The median and dispersion of the fragility curves in terms of peak ground
Table B1 – Fragility values in terms of PGA for two span continuous concrete box-
girder fragilities with diaphragm abutments.
Design Bridge class BSST-0 BSST-0 BSST-0 BSST-0 *
era
S-E1-S22-C-D 0.13 0.59 0.30 0.61 0.63 0.64 0.91 0.64 0.62
S-E1-S22-R-D 0.16 0.56 0.36 0.58 0.70 0.61 0.99 0.61 0.59
T-E1-S22-C-D 0.07 0.79 0.18 0.83 0.42 0.93 0.66 0.94 0.87
Era 11
T-E1-S22-R-D 0.09 0.68 0.18 0.70 0.35 0.74 0.50 0.74 0.71
M-E1-S22-C-D 0.07 0.74 0.16 0.75 0.36 1.00 0.53 1.00 0.87
M-E1-S22-R-D 0.06 0.81 0.15 0.91 0.37 1.40 0.59 1.39 1.13
S-E2-S22-C-D 0.15 0.61 0.53 0.60 1.12 0.66 1.45 0.66 0.63
S-E2-S22-O-D 0.17 0.61 0.63 0.59 1.68 0.62 2.20 0.61 0.61
T-E2-S22-C-D 0.11 0.68 0.38 0.64 0.89 0.74 1.17 0.75 0.70
Era 22
T-E2-S22-O-D 0.16 0.54 0.43 0.53 0.91 0.70 1.12 0.70 0.61
M-E2-S22-C-D 0.09 0.70 0.30 0.68 0.60 0.86 0.77 0.86 0.78
M-E2-S22-O-D 0.09 0.70 0.30 0.68 0.60 0.86 0.77 0.86 0.78
S-E3-S22-C-D 0.15 0.62 0.53 0.60 1.44 0.66 2.04 0.66 0.63
S-E3-S22-O-D 0.17 0.60 0.63 0.58 2.20 0.62 3.20 0.63 0.61
T-E3-S22-C-D 0.11 0.68 0.38 0.64 1.16 0.75 1.66 0.76 0.71
Era 33
T-E3-S22-O-D 0.17 0.53 0.43 0.54 1.12 0.69 1.49 0.69 0.61
M-E3-S22-C-D 0.09 0.70 0.30 0.69 0.77 0.87 1.07 0.87 0.78
M-E3-S22-O-D 0.09 0.70 0.30 0.69 0.77 0.87 1.08 0.87 0.78
207
Table B2 – Fragility values in terms of PGA for two span continuous concrete box-
girder fragilities with seat abutments.
Design Bridge class BSST-0 BSST-0 BSST-0 BSST-0 *
era
S-E1-S22-C-S 0.08 0.58 0.15 0.58 0.27 0.59 0.36 0.57 0.58
S-E1-S22-R-S 0.08 0.60 0.14 0.62 0.24 0.62 0.33 0.61 0.61
T-E1-S22-C-S 0.06 0.68 0.12 0.70 0.20 0.68 0.27 0.69 0.69
Era 11
T-E1-S22-R-S 0.09 0.71 0.14 0.71 0.23 0.69 0.31 0.69 0.70
M-E1-S22-C-S 0.05 0.64 0.09 0.64 0.19 0.62 0.27 0.60 0.62
M-E1-S22-R-S 0.08 0.65 0.13 0.65 0.23 0.64 0.31 0.63 0.64
S-E2-S22-C-S 0.10 0.64 0.47 0.64 0.89 0.73 1.20 0.73 0.69
S-E2-S22-O-S 0.12 0.55 0.49 0.53 0.96 0.60 1.29 0.60 0.57
T-E2-S22-C-S 0.08 0.63 0.33 0.60 0.64 0.67 0.87 0.68 0.64
Era 22
T-E2-S22-O-S 0.09 0.55 0.33 0.52 0.65 0.57 0.87 0.58 0.55
M-E2-S22-C-S 0.07 0.60 0.31 0.55 0.62 0.65 0.86 0.68 0.62
M-E2-S22-O-S 0.07 0.63 0.37 0.59 0.79 0.80 1.12 0.82 0.71
S-E3-S22-C-S 0.10 0.66 0.46 0.63 1.10 0.71 1.63 0.72 0.68
S-E3-S22-O-S 0.12 0.55 0.49 0.53 1.11 0.57 1.61 0.57 0.56
T-E3-S22-C-S 0.08 0.62 0.33 0.60 0.77 0.65 1.13 0.64 0.63
Era 33
T-E3-S22-O-S 0.09 0.54 0.33 0.52 0.72 0.56 1.01 0.56 0.54
M-E3-S22-C-S 0.07 0.58 0.32 0.57 0.75 0.63 1.12 0.63 0.60
M-E3-S22-O-S 0.07 0.65 0.37 0.60 0.94 0.76 1.46 0.75 0.69
208
Table B4 – Fragility values in terms of PGA for multi-span (S34) continuous
concrete box-girder fragilities with seat abutments.
Design Bridge class BSST-0 BSST-0 BSST-0 BSST-0 *
era
S-E1-S34-C-S 0.10 0.53 0.19 0.52 0.34 0.52 0.47 0.51 0.52
S-E1-S34-R-S 0.04 0.73 0.11 0.68 0.25 0.63 0.38 0.62 0.66
T-E1-S34-C-S 0.05 0.74 0.10 0.70 0.19 0.69 0.27 0.67 0.70
Era 11
T-E1-S34-R-S 0.01 1.00 0.04 0.82 0.14 0.73 0.25 0.69 0.81
M-E1-S34-C-S 0.01 0.93 0.04 0.84 0.12 0.77 0.19 0.75 0.82
M-E1-S34-R-S 0.01 0.93 0.04 0.84 0.12 0.77 0.19 0.75 0.82
S-E2-S34-C-S 0.08 0.73 0.42 0.75 0.80 0.97 1.15 0.98 0.86
S-E2-S34-O-S 0.11 0.63 0.53 0.63 1.11 0.80 1.61 0.82 0.72
T-E2-S34-C-S 0.03 0.79 0.29 0.68 0.63 1.02 1.01 1.09 0.89
Era 22
T-E2-S34-O-S 0.07 0.68 0.42 0.64 0.90 0.82 1.31 0.79 0.73
M-E2-S34-C-S 0.06 0.75 0.26 0.70 0.46 0.87 0.64 0.89 0.80
M-E2-S34-O-S 0.01 1.56 0.13 0.90 0.48 1.19 1.01 1.18 1.21
S-E3-S34-C-S 0.08 0.72 0.43 0.74 1.10 0.91 1.76 0.95 0.83
S-E3-S34-O-S 0.10 0.62 0.53 0.63 1.34 0.73 2.04 0.71 0.67
T-E3-S34-C-S 0.03 0.78 0.29 0.68 0.85 0.93 1.44 0.89 0.82
Era 33
T-E3-S34-O-S 0.07 0.69 0.41 0.64 1.06 0.69 1.58 0.67 0.67
M-E3-S34-C-S 0.06 0.74 0.26 0.70 0.61 0.85 0.93 0.85 0.78
M-E3-S34-O-S 0.01 1.50 0.13 0.90 0.80 1.03 1.42 0.83 1.06
209
APPENDIX C. COMPONENT FRAGILITY CURVES FOR BRIDGE
CLASSES
This appendix presents the component level fragility relationships for bridge
classes mentioned in Chapter 6. Table C1 and C2 documents the median and deviation
(logarithmic standard deviation) for the components for four damage states, for
diaphragam and seat abutment bridges, respectively. When the component median value
is more than 100, the corresponding median and dispersion values are reported as 99.00
and 0.00, respectively, to indicate that the contribution of the component to the system
bridges.
210
Ab-tran 0.22 0.63 0.82 0.63
T-E1-S22-C-D
Column 0.08 0.71 0.20 0.71 0.46 0.71 0.72 0.71
Deck-max 0.41 0.54 1.17 0.54 9.13 0.54 13.52 0.54
Fnd-tran 2.23 0.90 11.25 0.90
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.75 0.72 2.55 0.72
Ab-Act 0.35 0.73 0.98 0.73
Ab-tran 0.11 0.58 0.43 0.58
T-E1-S22-R-D
Column 0.11 0.75 0.21 0.75 0.43 0.75 0.62 0.75
Deck-max 0.52 0.62 1.40 0.62 9.56 0.62 13.80 0.62
Fnd-tran 2.98 1.09 23.60 1.09
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.83 0.65 2.27 0.65
Ab-Act 0.45 0.66 1.05 0.66
Ab-tran 0.15 0.63 0.52 0.63
M-E1-S22-C-D
Column 0.07 1.11 0.18 1.11 0.44 1.11 0.69 1.11
Deck-max 0.41 0.70 1.10 0.70 7.20 0.70 10.32 0.70
Fnd-tran 2.05 0.82 10.22 0.82
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.77 0.65 1.91 0.65
Ab-Act 0.43 0.65 0.93 0.65
Ab-tran 0.12 0.69 0.42 0.69
M-E1-S22-R-D
Column 0.06 1.36 0.15 1.36 0.41 1.36 0.69 1.36
Deck-max 0.55 0.60 1.54 0.60 11.35 0.60 16.63 0.60
Fnd-tran 1.75 1.08 10.24 1.08
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.99 0.90 3.31 0.90
Ab-Act 0.47 0.90 1.28 0.90
Ab-tran 0.15 0.65 0.58 0.65
S-E2-S22-C-D
Column 0.23 0.69 0.95 0.69 1.44 0.69 1.92 0.69
Deck-max 0.65 0.68 1.96 0.68 16.53 0.68 24.84 0.68
Fnd-tran 1.72 1.14 12.95 1.14
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.05 0.73 3.30 0.73
211
Ab-Act 0.50 0.74 1.36 0.74
Ab-tran 0.17 0.71 0.69 0.71
S-E2-S22-O-D
Column 0.30 0.73 1.48 0.73 2.36 0.73 3.24 0.73
Deck-max 0.73 0.63 2.52 0.63
Fnd-tran 1.03 1.00 7.89 1.00
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.94 0.78 3.35 0.78
Ab-Act 0.42 0.78 1.23 0.78
Ab-tran 0.17 0.64 0.79 0.64
T-E2-S22-C-D
Column 0.16 0.66 0.64 0.66 0.97 0.66 1.27 0.66
Deck-max 0.43 0.63 1.18 0.63
Fnd-tran 1.36 0.97 8.66 0.97
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.80 0.66 2.36 0.66
Ab-Act 0.40 0.65 1.00 0.65
Ab-tran 0.12 0.64 0.44 0.64
T-E2-S22-O-D
Column 0.25 0.57 0.83 0.57 1.17 0.57 1.49 0.57
Deck-max 0.56 0.58 1.62 0.58
Fnd-tran 1.12 1.05 10.61 1.05
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.52 0.96 8.03 0.96
Ab-Act 0.53 0.93 2.10 0.93
Ab-tran 0.16 0.59 0.59 0.59
M-E2-S22-C-D
Column 0.13 0.74 0.47 0.74 0.67 0.74 0.86 0.74
Deck-max 0.36 0.58 0.97 0.58
Fnd-tran 1.21 1.10 8.04 1.10
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.58 0.69 1.64 0.69
Ab-Act 0.31 0.69 0.73 0.69
Ab-tran 0.11 0.60 0.37 0.60
M-E2-S22-O-D
Column 0.01 13.40 0.02 13.40 0.01 13.40 0.38 13.40
Deck-max 0.61 0.50 1.78 0.50
Fnd-tran 0.59 0.74 3.18 0.74
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.79 0.58 2.35 0.58
212
Ab-Act 0.41 0.59 1.02 0.59
Ab-tran 0.16 0.50 0.63 0.50
S-E3-S22-C-D
Column 0.23 0.69 0.95 0.69 1.92 0.69 2.82 0.69
Deck-max 0.65 0.68 1.96 0.68
Fnd-tran 1.72 1.14 12.95 1.14
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.05 0.73 3.30 0.73
Ab-Act 0.50 0.74 1.36 0.74
Ab-tran 0.17 0.71 0.69 0.71
S-E3-S22-O-D
Column 0.30 0.73 1.48 0.73 3.24 0.73 4.99 0.73
Deck-max 0.73 0.63 2.52 0.63
Fnd-tran 1.03 1.00 7.89 1.00
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.94 0.78 3.35 0.78
Ab-Act 0.42 0.78 1.23 0.78
Ab-tran 0.17 0.64 0.79 0.64
T-E3-S22-C-D
Column 0.16 0.66 0.64 0.66 1.27 0.66 1.85 0.66
Deck-max 0.43 0.63 1.18 0.63
Fnd-tran 1.36 0.97 8.66 0.97
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.80 0.66 2.36 0.66
Ab-Act 0.40 0.65 1.00 0.65
Ab-tran 0.12 0.64 0.44 0.64
T-E3-S22-O-D
Column 0.25 0.57 0.83 0.57 1.49 0.57 2.06 0.57
Deck-max 0.56 0.58 1.62 0.58
Fnd-tran 1.12 1.05 10.61 1.05
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.52 0.96 8.03 0.96
Ab-Act 0.53 0.93 2.10 0.93
Ab-tran 0.16 0.59 0.59 0.59
M-E3-S22-C-D
Column 0.13 0.74 0.47 0.74 0.86 0.74 1.20 0.74
Deck-max 0.36 0.58 0.97 0.58
Fnd-tran 1.21 1.10 8.04 1.10
Fnd-rot 99.00 0.00 99.00 0.00
213
Ab-Pass 0.58 0.69 1.64 0.69
Ab-Act 0.31 0.69 0.73 0.69
Ab-tran 0.11 0.60 0.37 0.60
M-E3-S22-O-D
Column 0.01 13.40 0.02 13.40 0.01 13.40 0.38 13.40
Deck-max 0.61 0.50 1.78 0.50
Fnd-tran 0.59 0.74 3.18 0.74
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.79 0.58 2.35 0.58
Ab-Act 0.41 0.59 1.02 0.59
Ab-tran 0.16 0.50 0.63 0.50
S-E1-S34-C-D
Column 0.14 0.70 0.34 0.70 0.84 0.70 1.34 0.70
Deck-max 0.76 0.57 2.37 0.57
Fnd-tran 1.36 0.86 8.49 0.86
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.05 0.73 3.66 0.73
Ab-Act 0.46 0.68 1.29 0.68
Ab-tran 0.20 0.57 0.80 0.57
S-E1-S34-R-D
Column 0.06 1.32 0.24 1.32 0.99 1.32 2.03 1.32
Deck-max 1.06 0.58 3.47 0.58
Fnd-tran 1.09 0.80 6.96 0.80
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.09 0.62 3.71 0.62
Ab-Act 0.51 0.63 1.45 0.63
Ab-tran 0.29 0.60 1.30 0.60
T-E1-S34-C-D
Column 0.06 0.89 0.15 0.89 0.36 0.89 0.56 0.89
Deck-max 0.44 0.57 1.31 0.57
Fnd-tran 1.72 0.90 8.84 0.90
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.71 0.56 2.15 0.56
Ab-Act 0.35 0.57 0.90 0.57
Ab-tran 0.11 0.59 0.46 0.59
T-E1-S34-R-D
Column 0.07 0.91 0.17 0.91 0.41 0.91 0.66 0.91
Deck-max 0.54 0.53 1.57 0.53
Fnd-tran 3.86 0.93 37.46 0.93
214
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.86 0.63 2.58 0.63
Ab-Act 0.43 0.63 1.08 0.63
Ab-tran 0.15 0.55 0.58 0.55
M-E1-S34-C-D
Column 0.01 1.13 0.04 1.13 0.19 1.13 0.40 1.13
Deck-max 0.54 0.46 1.42 0.46
Fnd-tran 3.68 1.05 34.68 1.05
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.81 0.60 2.52 0.60
Ab-Act 0.41 0.62 1.08 0.62
Ab-tran 0.17 0.46 0.60 0.46
M-E1-S34-R-D
Column 0.01 1.13 0.04 1.13 0.19 1.13 0.40 1.13
Deck-max 0.54 0.46 1.42 0.46
Fnd-tran 3.68 1.05 34.68 1.05
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.81 0.60 2.52 0.60
Ab-Act 0.41 0.62 1.08 0.62
Ab-tran 0.17 0.46 0.60 0.46
S-E2-S34-C-D
Column 0.23 0.69 0.95 0.69 1.44 0.69 1.92 0.69
Deck-max 0.65 0.68 1.96 0.68
Fnd-tran 1.72 1.14 12.95 1.14
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.05 0.73 3.30 0.73
Ab-Act 0.50 0.74 1.36 0.74
Ab-tran 0.17 0.71 0.69 0.71
S-E2-S34-O-D
Column 0.30 0.73 1.48 0.73 2.36 0.73 3.24 0.73
Deck-max 0.73 0.63 2.52 0.63
Fnd-tran 1.03 1.00 7.89 1.00
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.94 0.78 3.35 0.78
Ab-Act 0.42 0.78 1.23 0.78
Ab-tran 0.17 0.64 0.79 0.64
T-E2-S34-C-D
Column 0.16 0.66 0.64 0.66 0.97 0.66 1.27 0.66
Deck-max 0.43 0.63 1.18 0.63
215
Fnd-tran 1.36 0.97 8.66 0.97
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.80 0.66 2.36 0.66
Ab-Act 0.40 0.65 1.00 0.65
Ab-tran 0.12 0.64 0.44 0.64
T-E2-S34-O-D
Column 0.25 0.57 0.83 0.57 1.17 0.57 1.49 0.57
Deck-max 0.56 0.58 1.62 0.58
Fnd-tran 1.12 1.05 10.61 1.05
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.52 0.96 8.03 0.96
Ab-Act 0.53 0.93 2.10 0.93
Ab-tran 0.16 0.59 0.59 0.59
M-E2-S34-C-D
Column 0.13 0.74 0.47 0.74 0.67 0.74 0.86 0.74
Deck-max 0.36 0.58 0.97 0.58
Fnd-tran 1.21 1.10 8.04 1.10
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.58 0.69 1.64 0.69
Ab-Act 0.31 0.69 0.73 0.69
Ab-tran 0.11 0.60 0.37 0.60
M-E2-S34-O-D
Column 0.13 0.74 0.47 0.74 0.67 0.74 0.86 0.74
Deck-max 0.36 0.58 0.97 0.58
Fnd-tran 1.21 1.10 8.04 1.10
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.58 0.69 1.64 0.69
Ab-Act 0.31 0.69 0.73 0.69
Ab-tran 0.11 0.60 0.37 0.60
S-E3-S34-C-D
Column 0.23 0.69 0.95 0.69 1.92 0.69 2.82 0.69
Deck-max 0.65 0.68 1.96 0.68
Fnd-tran 1.72 1.14 12.95 1.14
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.05 0.73 3.30 0.73
Ab-Act 0.50 0.74 1.36 0.74
Ab-tran 0.17 0.71 0.69 0.71
S-E3-S34-O-D
Column 0.30 0.73 1.48 0.73 3.24 0.73 4.99 0.73
216
Deck-max 0.73 0.63 2.52 0.63
Fnd-tran 1.03 1.00 7.89 1.00
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.94 0.78 3.35 0.78
Ab-Act 0.42 0.78 1.23 0.78
Ab-tran 0.17 0.64 0.79 0.64
T-E3-S34-C-D
Column 0.16 0.66 0.64 0.66 1.27 0.66 1.85 0.66
Deck-max 0.43 0.63 1.18 0.63
Fnd-tran 1.36 0.97 8.66 0.97
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.80 0.66 2.36 0.66
Ab-Act 0.40 0.65 1.00 0.65
Ab-tran 0.12 0.64 0.44 0.64
T-E3-S34-O-D
Column 0.25 0.57 0.83 0.57 1.49 0.57 2.06 0.57
Deck-max 0.56 0.58 1.62 0.58
Fnd-tran 1.12 1.05 10.61 1.05
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.52 0.96 8.03 0.96
Ab-Act 0.53 0.93 2.10 0.93
Ab-tran 0.16 0.59 0.59 0.59
M-E3-S34-C-D
Column 0.13 0.74 0.47 0.74 0.86 0.74 1.20 0.74
Deck-max 0.36 0.58 0.97 0.58
Fnd-tran 1.21 1.10 8.04 1.10
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.58 0.69 1.64 0.69
Ab-Act 0.31 0.69 0.73 0.69
Ab-tran 0.11 0.60 0.37 0.60
M-E3-S34-O-D
Column 0.13 0.74 0.47 0.74 0.86 0.74 1.20 0.74
Deck-max 0.36 0.58 0.97 0.58
Fnd-tran 1.21 1.10 8.04 1.10
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.58 0.69 1.64 0.69
Ab-Act 0.31 0.69 0.73 0.69
Ab-tran 0.11 0.60 0.37 0.60
217
Table C2 – Component level fragility relationships for seat abutment bridges.
218
Key 7.24 1.04 55.39 1.04
T-E1-S22-R-S
Column 0.12 0.69 0.23 0.69 0.45 0.69 0.63 0.69
Hinge 0.09 0.56 0.15 0.56 0.26 0.56 0.36 0.56
Deck-max 0.41 0.55 1.10 0.55
Fnd-tran 0.95 1.02 7.62 1.02
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.25 0.74 3.96 0.74
Ab-Act 0.57 0.80 1.38 0.80
Ab-tran 0.29 0.68 1.12 0.68
Bearing 0.14 0.56 0.43 0.56
Seal 0.26 0.56 4.50 0.56
Key 5.67 1.41 27.83 1.41
M-E1-S22-C-S
Column 0.04 1.19 0.11 1.19 0.29 1.19 0.48 1.19
Hinge 0.07 0.50 0.12 0.50 0.22 0.50 0.31 0.50
Deck-max 0.35 0.51 0.97 0.51
Fnd-tran 3.30 1.13 28.69 1.13
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.17 0.72 3.54 0.72
Ab-Act 0.60 0.71 1.48 0.71
Ab-tran 0.25 0.80 1.12 0.80
Bearing 0.11 0.48 0.36 0.48
Seal 0.22 0.50 4.83 0.50
Key 6.99 1.38 52.84 1.38
M-E1-S22-R-S
Column 0.13 0.71 0.25 0.71 0.49 0.71 0.70 0.71
Hinge 0.09 0.58 0.16 0.58 0.28 0.58 0.39 0.58
Deck-max 0.46 0.55 1.25 0.55
Fnd-tran 0.90 0.99 7.25 0.99
Fnd-rot 53.72 0.49 99.00 0.00
Ab-Pass 1.18 0.71 3.33 0.71
Ab-Act 0.62 0.77 1.41 0.77
Ab-tran 0.34 0.70 1.20 0.70
Bearing 0.16 0.58 0.47 0.58
Seal 0.28 0.58 4.89 0.58
Key 6.36 1.29 35.05 1.29
S-E2-S22-C-S
Column 0.16 0.66 0.66 0.66 1.01 0.66 1.34 0.66
219
Hinge 0.16 0.63 0.82 0.63 1.97 0.63 3.06 0.63
Deck-max 0.52 0.56 1.69 0.56
Fnd-tran 1.16 1.12 10.55 1.12
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.31 0.62 3.19 0.62
Ab-Act 0.73 0.64 1.48 0.64
Ab-tran 0.38 0.52 0.98 0.52
Bearing 0.13 0.59 0.76 0.59
Seal 0.34 0.63 17.71 0.63
Key 99.00 0.00 99.00 0.00
S-E2-S22-O-S
Column 0.20 0.60 0.86 0.60 1.32 0.60 1.77 0.60
Hinge 0.15 0.54 0.66 0.54 1.44 0.54 2.14 0.54
Deck-max 0.50 0.52 1.54 0.52
Fnd-tran 0.75 0.98 5.06 0.98
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.24 0.59 3.17 0.59
Ab-Act 0.66 0.60 1.35 0.60
Ab-tran 0.39 0.53 1.09 0.53
Bearing 0.14 0.54 0.70 0.54
Seal 0.30 0.54 10.35 0.54
Key 99.00 0.00 99.00 0.00
T-E2-S22-C-S
Column 0.11 0.59 0.53 0.59 0.84 0.59 1.14 0.59
Hinge 0.12 0.56 0.52 0.56 1.14 0.56 1.70 0.56
Deck-max 0.36 0.54 1.11 0.54
Fnd-tran 1.32 0.94 9.35 0.94
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.02 0.53 2.65 0.53
Ab-Act 0.55 0.60 1.15 0.60
Ab-tran 0.25 0.49 0.71 0.49
Bearing 0.10 0.54 0.50 0.54
Seal 0.23 0.56 8.27 0.56
Key 99.00 0.00 99.00 0.00
T-E2-S22-O-S
Column 0.14 0.65 0.61 0.65 0.92 0.65 1.23 0.65
Hinge 0.12 0.46 0.48 0.46 1.02 0.46 1.50 0.46
Deck-max 0.35 0.44 1.02 0.44
Fnd-tran 0.47 0.89 2.69 0.89
Fnd-rot 99.00 0.00 99.00 0.00
220
Ab-Pass 1.00 0.51 2.46 0.51
Ab-Act 0.46 0.42 0.90 0.42
Ab-tran 0.29 0.46 0.80 0.46
Bearing 0.10 0.46 0.47 0.46
Seal 0.22 0.46 6.89 0.46
Key 99.00 0.00 99.00 0.00
M-E2-S22-C-S
Column 0.11 0.67 0.50 0.67 0.79 0.67 1.07 0.67
Hinge 0.11 0.53 0.50 0.53 1.16 0.53 1.76 0.53
Deck-max 0.35 0.47 1.07 0.47
Fnd-tran 0.97 0.90 6.15 0.90
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.94 0.42 2.20 0.42
Ab-Act 0.54 0.42 1.06 0.42
Ab-tran 0.27 0.41 0.67 0.41
Bearing 0.09 0.52 0.49 0.52
Seal 0.22 0.53 9.46 0.53
Key 99.00 0.00 99.00 0.00
M-E2-S22-O-S
Column 0.14 0.75 0.66 0.75 1.04 0.75 1.41 0.75
Hinge 0.12 0.54 0.58 0.54 1.35 0.54 2.07 0.54
Deck-max 0.40 0.50 1.27 0.50
Fnd-tran 0.49 0.87 3.89 0.87
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.10 0.52 2.74 0.52
Ab-Act 0.63 0.52 1.28 0.52
Ab-tran 0.32 0.42 0.81 0.42
Bearing 0.10 0.58 0.58 0.58
Seal 0.25 0.54 11.29 0.54
Key 99.00 0.00 99.00 0.00
S-E3-S22-C-S
Column 0.16 0.66 0.66 0.66 1.34 0.66 1.99 0.66
Hinge 0.16 0.63 0.82 0.63 1.97 0.63 3.06 0.63
Deck-max 0.52 0.56 1.69 0.56
Fnd-tran 1.16 1.12 10.55 1.12
Fnd-rot 99.00 1.15 99.00 1.15
Ab-Pass 1.31 0.62 3.19 0.62
Ab-Act 0.73 0.64 1.48 0.64
Ab-tran 0.38 0.52 0.98 0.52
221
Bearing 0.13 0.59 0.76 0.59
Seal 0.34 0.63 17.71 0.63
Key 99.00 0.00 99.00 0.00
S-E3-S22-O-S
Column 0.20 0.60 0.86 0.60 1.77 0.60 2.64 0.60
Hinge 0.15 0.54 0.66 0.54 1.44 0.54 2.14 0.54
Deck-max 0.50 0.52 1.54 0.52
Fnd-tran 0.75 0.98 5.06 0.98
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.24 0.59 3.17 0.59
Ab-Act 0.66 0.60 1.35 0.60
Ab-tran 0.39 0.53 1.09 0.53
Bearing 0.14 0.54 0.70 0.54
Seal 0.30 0.54 10.35 0.54
Key 99.00 0.00 99.00 0.00
T-E3-S22-C-S
Column 0.11 0.59 0.53 0.59 1.14 0.59 1.75 0.59
Hinge 0.12 0.56 0.52 0.56 1.14 0.56 1.70 0.56
Deck-max 0.36 0.54 1.11 0.54
Fnd-tran 1.32 0.94 9.35 0.94
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.02 0.53 2.65 0.53
Ab-Act 0.55 0.60 1.15 0.60
Ab-tran 0.25 0.49 0.71 0.49
Bearing 0.10 0.54 0.50 0.54
Seal 0.23 0.56 8.27 0.56
Key 99.00 0.00 99.00 0.00
T-E3-S22-O-S
Column 0.14 0.65 0.61 0.65 1.23 0.65 1.82 0.65
Hinge 0.12 0.46 0.48 0.46 1.02 0.46 1.50 0.46
Deck-max 0.35 0.44 1.02 0.44
Fnd-tran 0.47 0.89 2.69 0.89
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.00 0.51 2.46 0.51
Ab-Act 0.46 0.42 0.90 0.42
Ab-tran 0.29 0.46 0.80 0.46
Bearing 0.10 0.46 0.47 0.46
Seal 0.22 0.46 6.89 0.46
Key 99.00 0.00 99.00 0.00
222
M-E3-S22-C-S
Column 0.11 0.67 0.50 0.67 1.07 0.67 1.63 0.67
Hinge 0.11 0.53 0.50 0.53 1.16 0.53 1.76 0.53
Deck-max 0.35 0.47 1.07 0.47
Fnd-tran 0.97 0.90 6.15 0.90
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 0.94 0.42 2.20 0.42
Ab-Act 0.54 0.42 1.06 0.42
Ab-tran 0.27 0.41 0.67 0.41
Bearing 0.09 0.52 0.49 0.52
Seal 0.22 0.53 9.46 0.53
Key 99.00 0.00 99.00 0.00
M-E3-S22-O-S
Column 0.14 0.75 0.66 0.75 1.41 0.75 2.15 0.75
Hinge 0.12 0.54 0.58 0.54 1.35 0.54 2.07 0.54
Deck-max 0.40 0.50 1.27 0.50
Fnd-tran 0.49 0.87 3.89 0.87
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.10 0.52 2.74 0.52
Ab-Act 0.63 0.52 1.28 0.52
Ab-tran 0.32 0.42 0.81 0.42
Bearing 0.10 0.58 0.58 0.58
Seal 0.25 0.54 11.29 0.54
Key 99.00 0.00 99.00 0.00
S-E1-S34-C-S
Column 0.12 0.58 0.28 0.58 0.63 0.58 0.96 0.58
Hinge 0.13 0.50 0.22 0.50 0.40 0.50 0.56 0.50
Deck-max 0.68 0.48 1.99 0.48
Fnd-tran 1.24 0.81 7.07 0.81
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.47 0.72 4.39 0.72
Ab-Act 0.74 0.73 1.78 0.73
Ab-tran 0.34 0.67 1.38 0.67
Bearing 0.22 0.50 0.71 0.50
Seal 0.40 0.50 8.12 0.50
Key 3.98 1.04 18.68 1.04
S-E1-S34-R-S
Column 0.06 1.07 0.17 1.07 0.50 1.07 0.88 1.07
Hinge 0.10 0.49 0.18 0.49 0.33 0.49 0.47 0.49
223
Deck-max 0.61 0.51 1.82 0.51
Fnd-tran 1.59 0.95 12.69 0.95
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.37 0.67 3.86 0.67
Ab-Act 0.71 0.67 1.62 0.67
Ab-tran 0.54 0.88 2.80 0.88
Bearing 0.18 0.49 0.61 0.49
Seal 0.33 0.49 8.00 0.49
Key 5.70 1.14 32.67 1.14
T-E1-S34-C-S
Column 0.06 0.90 0.13 0.90 0.29 0.90 0.44 0.90
Hinge 0.09 0.51 0.16 0.51 0.27 0.51 0.37 0.51
Deck-max 0.39 0.58 1.01 0.58
Fnd-tran 2.93 0.87 18.56 0.87
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.06 0.64 3.31 0.64
Ab-Act 0.54 0.78 1.28 0.78
Ab-tran 0.24 0.78 0.95 0.78
Bearing 0.15 0.52 0.44 0.52
Seal 0.27 0.51 4.34 0.51
Key 3.12 0.99 12.03 0.99
T-E1-S34-R-S
Column 0.02 1.10 0.08 1.10 0.26 1.10 0.49 1.10
Hinge 0.08 0.56 0.14 0.56 0.27 0.56 0.39 0.56
Deck-max 0.45 0.54 1.33 0.54
Fnd-tran 5.04 1.29 86.51 1.29
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.09 0.60 3.18 0.60
Ab-Act 0.54 0.65 1.29 0.65
Ab-tran 0.29 0.70 1.14 0.70
Bearing 0.13 0.56 0.47 0.56
Seal 0.27 0.56 7.18 0.56
Key 24.57 1.96 99.00 1.96
M-E1-S34-C-S
Column 0.01 1.36 0.02 1.36 0.11 1.36 0.24 1.36
Hinge 0.11 0.58 0.18 0.58 0.32 0.58 0.44 0.58
Deck-max 0.45 0.45 1.23 0.45
Fnd-tran 3.54 1.30 28.44 1.30
Fnd-rot 99.00 0.00 99.00 0.00
224
Ab-Pass 1.55 1.01 4.35 1.01
Ab-Act 0.75 0.99 1.67 0.99
Ab-tran 0.22 0.60 0.71 0.60
Bearing 0.16 0.54 0.50 0.54
Seal 0.32 0.58 5.58 0.58
Key 15.07 1.94 99.00 1.94
M-E1-S34-R-S
Column 0.01 1.36 0.02 1.36 0.11 1.36 0.24 1.36
Hinge 0.11 0.58 0.18 0.58 0.32 0.58 0.44 0.58
Deck-max 0.45 0.45 1.23 0.45
Fnd-tran 3.54 1.30 28.44 1.30
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.55 1.01 4.35 1.01
Ab-Act 0.75 0.99 1.67 0.99
Ab-tran 0.22 0.60 0.71 0.60
Bearing 0.16 0.54 0.50 0.54
Seal 0.32 0.58 5.58 0.58
Key 15.07 1.94 99.00 1.94
S-E2-S34-C-S
Column 0.11 0.82 0.59 0.82 0.97 0.82 1.36 0.82
Hinge 0.22 0.64 1.12 0.64 2.67 0.64 4.16 0.64
Deck-max 0.65 0.57 2.11 0.57
Fnd-tran 1.46 1.02 12.66 1.02
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.75 0.61 4.55 0.61
Ab-Act 0.99 0.63 2.11 0.63
Ab-tran 0.39 0.48 1.05 0.48
Bearing 0.17 0.61 0.96 0.61
Seal 0.46 0.64 23.96 0.64
Key 99.00 2.32 99.00 2.32
S-E2-S34-O-S
Column 0.15 1.05 1.05 1.05 1.85 1.05 2.70 1.05
Hinge 0.21 0.56 1.00 0.56 2.29 0.56 3.50 0.56
Deck-max 0.65 0.55 2.22 0.55
Fnd-tran 0.87 0.90 5.98 0.90
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.43 0.53 3.47 0.53
Ab-Act 0.81 0.54 1.64 0.54
Ab-tran 0.50 0.60 1.43 0.60
225
Bearing 0.17 0.59 1.01 0.59
Seal 0.43 0.56 18.68 0.56
Key 99.00 0.00 99.00 0.00
T-E2-S34-C-S
Column 0.03 1.38 0.32 1.38 0.66 1.38 1.09 1.38
Hinge 0.14 0.54 0.63 0.54 1.42 0.54 2.14 0.54
Deck-max 0.40 0.49 1.24 0.49
Fnd-tran 1.55 0.92 11.68 0.92
Fnd-rot 99.00 0.67 99.00 0.67
Ab-Pass 1.08 0.55 2.71 0.55
Ab-Act 0.58 0.56 1.20 0.56
Ab-tran 0.29 0.48 0.83 0.48
Bearing 0.11 0.49 0.57 0.49
Seal 0.28 0.54 11.00 0.54
Key 99.00 0.00 99.00 0.00
T-E2-S34-O-S
Column 0.05 1.22 0.65 1.22 1.35 1.22 2.22 1.22
Hinge 0.14 0.54 0.69 0.54 1.59 0.54 2.43 0.54
Deck-max 0.45 0.55 1.48 0.55
Fnd-tran 0.65 0.77 3.91 0.77
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.32 0.69 3.29 0.69
Ab-Act 0.68 0.69 1.35 0.69
Ab-tran 0.34 0.50 0.87 0.50
Bearing 0.11 0.58 0.67 0.58
Seal 0.29 0.54 13.16 0.54
Key 99.00 0.00 99.00 0.00
M-E2-S34-C-S
Column 0.07 0.96 0.36 0.96 0.58 0.96 0.80 0.96
Hinge 0.15 0.55 0.69 0.55 1.53 0.55 2.31 0.55
Deck-max 0.44 0.52 1.40 0.52
Fnd-tran 1.11 0.81 7.42 0.81
Fnd-rot 99.00 0.70 99.00 0.70
Ab-Pass 1.12 0.55 2.87 0.55
Ab-Act 0.62 0.60 1.28 0.60
Ab-tran 0.30 0.50 0.82 0.50
Bearing 0.12 0.58 0.66 0.58
Seal 0.30 0.55 11.72 0.55
Key 99.00 0.00 99.00 0.00
226
M-E2-S34-O-S
Column 0.01 1.17 0.31 1.17 0.89 1.17 1.82 1.17
Hinge 0.15 0.55 0.66 0.55 1.44 0.55 2.14 0.55
Deck-max 0.52 0.56 1.63 0.56
Fnd-tran 0.89 0.95 8.41 0.95
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.08 0.54 2.59 0.54
Ab-Act 0.61 0.55 1.22 0.55
Ab-tran 0.40 0.50 1.11 0.50
Bearing 0.14 0.56 0.69 0.56
Seal 0.30 0.55 10.25 0.55
Key 99.00 0.00 99.00 0.00
S-E3-S34-C-S
Column 0.11 0.82 0.59 0.82 1.36 0.82 2.16 0.82
Hinge 0.22 0.64 1.12 0.64 2.67 0.64 4.16 0.64
Deck-max 0.65 0.57 2.11 0.57
Fnd-tran 1.46 1.02 12.66 1.02
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.75 0.61 4.55 0.61
Ab-Act 0.99 0.63 2.11 0.63
Ab-tran 0.39 0.48 1.05 0.48
Bearing 0.17 0.61 0.96 0.61
Seal 0.46 0.64 23.96 0.64
Key 99.00 0.00 99.00 0.00
S-E3-S34-O-S
Column 0.15 1.05 1.05 1.05 2.70 1.05 4.55 1.05
Hinge 0.21 0.56 1.00 0.56 2.29 0.56 3.50 0.56
Deck-max 0.65 0.55 2.22 0.55
Fnd-tran 0.87 0.90 5.98 0.90
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.43 0.53 3.47 0.53
Ab-Act 0.81 0.54 1.64 0.54
Ab-tran 0.50 0.60 1.43 0.60
Bearing 0.17 0.59 1.01 0.59
Seal 0.43 0.56 18.68 0.56
Key 99.00 0.00 99.00 0.00
T-E3-S34-C-S
Column 0.03 1.38 0.32 1.38 1.09 1.38 2.15 1.38
227
Hinge 0.14 0.54 0.63 0.54 1.42 0.54 2.14 0.54
Deck-max 0.40 0.49 1.24 0.49
Fnd-tran 1.55 0.92 11.68 0.92
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.08 0.55 2.71 0.55
Ab-Act 0.58 0.56 1.20 0.56
Ab-tran 0.29 0.48 0.83 0.48
Bearing 0.11 0.49 0.57 0.49
Seal 0.28 0.54 11.00 0.54
Key 99.00 0.00 99.00 0.00
T-E3-S34-O-S
Column 0.05 1.22 0.65 1.22 2.22 1.22 4.35 1.22
Hinge 0.14 0.54 0.69 0.54 1.59 0.54 2.43 0.54
Deck-max 0.45 0.55 1.48 0.55
Fnd-tran 0.65 0.77 3.91 0.77
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.32 0.69 3.29 0.69
Ab-Act 0.68 0.69 1.35 0.69
Ab-tran 0.34 0.50 0.87 0.50
Bearing 0.11 0.58 0.67 0.58
Seal 0.29 0.54 13.16 0.54
Key 99.00 0.00 99.00 0.00
M-E3-S34-C-S
Column 0.07 0.96 0.36 0.96 0.80 0.96 1.24 0.96
Hinge 0.15 0.55 0.69 0.55 1.53 0.55 2.31 0.55
Deck-max 0.44 0.52 1.40 0.52
Fnd-tran 1.11 0.81 7.42 0.81
Fnd-rot 99.00 0.00 99.00 0.00
Ab-Pass 1.12 0.55 2.87 0.55
Ab-Act 0.62 0.60 1.28 0.60
Ab-tran 0.30 0.50 0.82 0.50
Bearing 0.12 0.58 0.66 0.58
Seal 0.30 0.55 11.72 0.55
Key 99.00 0.00 99.00 0.00
M-E3-S34-O-S
Column 0.01 1.17 0.31 1.17 1.82 1.17 4.85 1.17
Hinge 0.15 0.55 0.66 0.55 1.44 0.55 2.14 0.55
Deck-max 0.52 0.56 1.63 0.56
Fnd-tran 0.89 0.95 8.41 0.95
228
Fnd-rot 99.00 0.51 99.00 0.51
Ab-Pass 1.08 0.54 2.59 0.54
Ab-Act 0.61 0.55 1.22 0.55
Ab-tran 0.40 0.50 1.11 0.50
Bearing 0.14 0.56 0.69 0.56
Seal 0.30 0.55 10.25 0.55
Key 99.00 1.79 99.00 1.79
229
REFERENCES
Ala Saadeghvaziri, M., and Rashidi, S. (1998). Effect of steel bearings on seismic response of
bridges in Eastern United States. Proceedings of 6th US National Conf. on Earthquake
Engineering. Oakland, California
ATC. (1985). Earthquake damage evaluation data for California, Report No. ATC-6. Retrieved
from Applied Technology Council, Redwood City, California.
Avsar, O., Yakut A and Caner, A. (2001). Analytical Fragility Curves for Ordinary Highway
Bridges in Turkey. Earthquake Spectra, 27(4), pp: 971-996.
Baker, J. W., Lin, T., Shahi, S. K., and Jayaram, N. (2011). New ground motion selection
procedures and selected motions for the PEER transportation research program. Pacific
Earthquake Engineering Research Center.
Banerjee, S., and Shinozuka, M. (2007). Nonlinear static procedure for seismic vulnerability
assessment of bridges. Computer‐Aided Civil and Infrastructure Engineering, 22(4),
pp:293-305.
Banerjee, S., and Shinozuka, M. (2008). Mechanistic quantification of RC bridge damage states
under earthquake through fragility analysis. Probabilistic Engineering Mechanics, 23,
pp:12-22.
Basöz, N., and Mander, J. (1999). Enhancement of the highway transportation lifeline module in
HAZUS. National Institute of Building Sciences, 16(1), pp:31-40.
Basoz, N., and Kiremidjian, A. S. (1998). Evaluation of bridge damage data from the Loma
Prieta and Northridge, California earthquakes. California earthquakes (No. MCEER-
98-0004).
BDS (1990). Bridge design details, Caifornia Department of Trnasportations, Sacramento. CA.
BDA (2012). Bridge design aids, Caifornia Department of Trnasportations, Sacramento. CA.
Başöz, N., and Kiremidjian, A. S. (1996). Risk assessment for highway transportation systems:
John A. Blume Earthquake Engineering Center.
Bavirisetty, R., Vinayagamoorthy, M., and Duan, L. (2003). Dynamic Analysis, Bridge
Engineering – Seismic Design, Edited by Wai-Fah Chen and Lian Duan, CRC Press LLC,
Boca Raton, FL, ISBN: 0-8493-1683-9/02.
Bradley, B. A. (2010). Epistemic Uncertainties in Component Fragility Functions. Earthquake
Spectra 26(1), pp: 41-62.
Calderone, A., Lehman, D. E., and Moehle J. P. (2001). Behavior of reinforced concrete bridge
columns having varying aspect ratios and varying lengths of confinement, Pacific
Earthquake Engineering Research Center.
Caltrans (2017). Personal communication with the P266, Task 1780 Fragility project panel
members including Roblee C, Shantz T, Turner L. California Department of
Transportation. Sacramento.
Caltrans, S. (2010). Caltrans Seismic Design Criteria version 1.6. California Department of
Transportation, Sacramento, California.
Celik, O. C., and B. R. Ellingwood (2010). Seismic fragilities for non-ductile reinforced concrete
frames - Role of aleatoric and epistemic uncertainties. Structural Safety 32(1), pp: 1-12.
Chai, Y. H., Priestley, M. N., and Seible, F. (1991). Seismic retrofit of circular bridge columns
for enhanced flexural performance. ACI Structural Journal, 88.
230
Chang, G. A. and Mander, J. B. (1994). Seismic energy based fatigue damage analysis of bridge
columns: Part 1: Evaluation of seismic capacity. In Technical Report, US National
Center for Earthquake Engineering Research.
Cheok, G. S. and Stone, W. C. (1990). Behavior of 1/6-scale model bridge columns subjected to
inelastic cyclic loading. ACI Structural Journal 87(6).
Choi, E. (2002). Seismic analysis and retrofit of mid-America bridges. Ph.D Thesis, Georgia
Institute of Technology, Atlanta, USA.
Coffman, H. L., Marsh, M. L., and Brown, C. B. (1993). Seismic durability of retrofitted
reinforced-concrete columns. Journal of Structural Engineering, 119(5), 1643-1661.
Cornell, C., Jalayer, F., Hamburger, R., and Foutch, D. (2002). Probabilistic Basis for 2000 SAC
Federal Emergency Management Agency Steel Moment Frame Guidelines. Journal of
Structural Engineering, 128(4), pp:526-533.
Cornell, C. A., and Krawinkler, H. (2000). Progress and challenges in seismic performance
assessment. PEER Center News, 3(2), 1-3.
Correal, J. F., Saiidi, M. S., Sanders, D., and El-Azazy, S. (2007). Analytical evaluation of bridge
columns with double interlocking spirals. ACI structural journal, 104(3), 314.
Dicleli, M., and Bruneau, M. (1995). Seismic performance of single-span simply supported and
continuous slab-on-girder steel highway bridges. Journal of structural engineering,
121(10), pp:1497-1506.
Dukes, J. D. (2013). Application of bridge specific fragility analysis in the seismic design process
of bridges in california. Ph.D Thesis, Georgia Institute of Technology, Atlanta, USA.
Dukes, J., Mangalathu, S., Padgett, J. E., and DesRoches, R. (2017). Development of bridge-
specific fragility methedology to improve the seismic resilience of bridges, Earthquakes
and Structures (in review).
Esmaeily, A., and Xiao, Y. (2005). Behavior of reinforced concrete columns under variable axial
loads: analysis. ACI Structural Journal, 102(5), 736.
Fang, J. Q., Li, Q. S., Jeary, A. P., and Liu, D. K. (1999). Damping of tall buildings: its evaluation
and probabilistic characteristics. The Structural Design of Tall Buildings, 8(2), 145-153.
Friedman, J., Hastie, T., and Tibshirani, R. (2001) The elements of statistical learning. Springer
series in statistics Springer, Berlin.
Filippou, F. C., Popov, E. P., and Bertero, B. V. (1983). Effects of bond deterioration on
hysteretic behavior of reinforced concrete joints, Report No. FEMA-351. Washington
DC: SAC Joint Venture.
Fajfar, P. (2000). A nonlinear analysis method for performance-based seismic design. Earthquake
Spectra, 16(3), 573-592.
Gardoni, P., Mosalam, K. M., and der Kiureghian, A. (2003). Probabilistic seismic demand
models and fragility estimates for RC bridges. Journal of Earthquake Engineering, 7,
pp:79-106.
Ghosh, J., Padgett, J. E., and Dueñas-Osorio, L. (2013). Surrogate modeling and failure surface
visualization for efficient seismic vulnerability assessment of highway bridges.
Probabilistic Engineering Mechanics, 34, pp:189-199.
231
Haselton, C., A. Whittaker, A. Hortacsu, J. Baker, J. Bray and Grant, D. (2012). Selecting and
scaling earthquake ground motions for performing response-history analyses.
Proceedings of the 15th World Conference on Earthquake Engineering.
HAZUS-MH. (2003). Multi-Hazard Loss Estimation Methodology: Earthquake Model.
Department of Homeland Security, FEMA, Washington, DC.
Hesterberg, T., Choi, N. H., Meier, L., Fraley, C., (2008). Least angle and ℓ1 penalized
regression: A review. Statistics Surveys, 2,pp:61-93.
Hoerl, A. E., and Kennar., R. W. (1970). Ridge regression: Biased estimation for nonorthogonal
problems, Technometrics, 12,pp:55-67.
Hose, Y. D., Priestley, M. and Seible, F. (1997). Strategic relocation of plastic hinges in bridge
columns. Structural Systems Research Project, 97/05, University of California, San
Diego.
Huitema, B. E. (1990). Analysis of Covariance, in Encyclopedia of Statistics in Behavioral
Science. John Wiley and Sons, Ltd: Hoboken, New Jersey.
Hwang, H., Jernigan, J. B., and Lin, Y.-W. (2000). Evaluation of seismic damage to Memphis
bridges and highway systems. Journal of Bridge Engineering, 5(4), pp:322-330.
Jangid, R. S. (2004). Seismic response of isolated bridges. Journal of Bridge Engineering, 9(2),
pp:156–166.
Jaradat, O. A., McLean, D. I. and Marsh, M. L. (1998). Performance of Existing Bridge Columns
under Cyclic Loading Part 1: Experimental Results and Observed Behavior. ACI
Structural Journal 95(6).
Jeon, J. S. (2013). Aftershock vulnerability assessment of damaged reinforced concrete buildings
in California, PhD. Thesis, Georgia Tech, Atlanta.
Jeon, J. S., Shafieezadeh, A., Lee, D. H., Choi, E., and DesRoches, R. (2015). Damage
assessment of older highway bridges subjected to three-dimensional ground motions:
characterization of shear–axial force interaction on seismic fragilities. Engineering
Structures, 87, pp: 47-57.
Jeon, J.-S., Mangalathu, S., Song, J., and DesRoches, R. (2017). Parameterized seismic fragility
curves for curved multi-frame concrete box-girder bridges using Bayesian parameter
estimation. Journal of Earthquake Engineering (In press).
Jeong, S.-H., and Elnashai, A. S. (2007). Probabilistic fragility analysis parameterized by
fundamental response quantities. Engineering Structures, 29(6), pp:1238-1251.
Ji, J., Elnashai, A. S., and Kuchma, D. A. (2007). An analytical framework for seismic fragility
analysis of RC high-rise buildings. Engineering structures, 29(12), pp:3197-3209.
Kameshwar, S. and Padgett, J. E. (2014) Multi-hazard risk assessment of highway bridges
subjected to earthquake and hurricane hazards. Engineering Structures, 78, pp:154-166.
Keselman, H., Huberty, C. J., Lix, L. M., Olejnik, S., Cribbie, R. A., Donahue, B., Kowalchuk,
R. K., Lowman, L. L., Petoskey, M. D., and Keselman, J. C. (1998). Statistical practices
of educational researchers: An analysis of their ANOVA, MANOVA, and ANCOVA
analyses. Review of Educational Research, 68, pp:350-386.
Kim, S.-H., and Shinozuka, M. (2004). Development of fragility curves of bridges retrofitted by
column jacketing. Probabilistic Engineering Mechanics, 19(1), pp:105-112.
232
Kircher, C. A., Whitman, R. V., and Holmes, W. T. (2006). HAZUS earthquake loss estimation
methods. Natural Hazards Review, 7(2), pp:45-59.
Kowalsky, M. J. and Priestley, M. N. (2000). Improved analytical model for shear strength of
circular reinforced concrete columns in seismic regions. ACI Structural Journal 97(3).
Kolmogorov, A. N. (1933). Sulla determinazione empirica di una legge di distribuzione.
Kruskal, W. H., and Wallis, W. A. (1952). Use of ranks in one-criterion variance analysis.
Journal of the American statistical Association, 47(260),pp:583-621.
Kunnath, S. K., El-Bahy, A.,Taylor, A. W. and Stone, W. C. (1997). Cumulative seismic damage
of reinforced concrete bridge piers. Technical Report NCEER, US National Center for
Earthquake Engineering Research.
Kunnath, S. K., Larson, L., and Miranda, E. (2006) Modelling considerations in probabilistic
performance‐based seismic evaluation: case study of the I‐880 viaduct. Earthquake
Engineering and Structural Dynamics, 35, pp:57-75.
Kvam, P. H., and Vidakovic, B. (2007). Nonparametric statistics with applications to science
and engineering. John Wiley and Sons..
Lehman, D., Moehle, J., Mahin, S., Calderone, A. and Henry. L. (2004). Experimental
evaluation of the seismic performance of reinforced concrete bridge columns. Journal of
Structural Engineering, 130, pp:869-879.
Lehman, D. E. and Moehle, J. P. (2000). Seismic performance of well-confined concrete bridge
columns, Pacific Earthquake Engineering Research Center.
Mackie, K., and Stojadinovic, B. (2001). Probabilistic seismic demand model for California
highway bridges. Journal of Bridge Engineering, 6(6), 468-481.
Mackie, K. R., and Stojadinović, B. (2005). Fragility basis for California highway overpass
bridge seismic decision making. Pacific Earthquake Engineering Research Center,
College of Engineering, University of California, Berkeley.
Mackie, K. R., Cronin, K. J., Nielson, B. G. (2011). Response Sensitivity of Highway Bridges to
Randomly Oriented Multi-Component Earthquake Excitation, Journal of Earthquake
Engineering, 15(6), pp: 850-876.
Mander, J., Priestley, M., and Park, R. (1988). Theoretical Stress‐Strain Model for Confined
Concrete. Journal of Structural Engineering, 114(8), pp:1804-1826.
Mander, J. B., and Basöz, N. (1999). Seismic fragility curve theory for highway bridges. Paper
presented at the Optimizing post-earthquake lifeline system reliability.
Mangalathu, S., Jeon, J.-S., Soleimani, F., DesRoches, R., Padgett, J., and Jiang, J. (2015a).
Seismic vulnerability of multi-span bridges: An analytical perspective. Proceedings of
the 10th Pacific Conference on Earthquake Engineering, Sydney, Australia.
Mangalathu, S., Jeon, J.-S., DesRoches, R., and Padgett, J. (2015b). Analysis of Covariance to
Capture the Importance of Bridge Attributes on the Probabilistic Seismic Demand
Model. Proceedings of the 10th Pacific Conference on Earthquake Engineering, Sydney,
Australia.
Mangalathu, S., Jeon, J.-S., DesRoches, R., and Padgett, J. (2016a). ANCOVA-based grouping
of bridge classes for seismic fragility assessment. Engineering Structures, 123, 379-394.
233
Mangalathu, S., Jeon J-S., DesRoches, R., and Padgett, J. E. (2016b). Application of Bayesian
Methods to Probabilistic Demand Analyses of Concrete Box-Girder Bridges.
Proceedings of the Geotechnical and Structural Engineering Congress, Phoenix, Arizona.
Mangalathu, S., Jeon, J.-S., DesRoches, R., and Padgett, J. (2017a). Performance-based grouping
methods of bridge classes for regional seismic risk assessment: Application of ANOVA,
ANCOVA, and non-parameteric approach. Earthquake Engineering and Structural
Dynamics (In review).
Mangalathu, S., Soleimani, F., DesRoches, R., and Padgett, J. E. (2017b). ANOVA based
grouping of bridge classes. Proceedings of the 16th World Conference on Earthquake
Engineering, Santiago, Chile.
Mangalathu, S., Jeon J-S., Soleimani, F., DesRoches, R., Padgett, J. E., and Jiang, J. (2017c).
Sensitivity of Fragility Curves to Parameter Uncertainty using Lasso. Proceedings of the
16th World Conference on Earthquake Engineering, Santiago, Chile.
Mangalathu, S., Soleimani, F., DesRoches, R., and Padgett, J. E. (2017d). Bridge Classes for
Regional Risk assessment of Box-Girder Bridges in California: Improving HAZUS
Model. Proceedings of the Structures Congress, Denver, Colorado.
Mangalathu, S., Jeon J-S., DesRoches, R., (2017e) Identification of critical parameters on the
seismic performance of concrete bridges using Lasso regression, Earthquake
Engineering and Structural Dynamics (In review).
Mazzoni, S, McKenna. F., Scott, M. H., Fenves, G. L. (2006) OpenSees command language
manual. Pacific Earthquake Engineering Research (PEER) Center.
Megally, S. H., Silva, F. P., and Seible, F. (2002). Seismic Response of Sacrificial Shear Keys in
Bridge Abutments. Department of Structural Engineering, University of California, San
Diego.
Mehr, M. and Zaghi, A. E. (2016). Seismic response of multi-frame bridges. Bulletin of
Earthquake Engineering. 14, pp:1219-43.
Moehle, J., and Deierlein, G. G. (2004). A framework methodology for performance-based
earthquake engineering. Proceedings of the 13th World Conference on Earthquake
Engineering.
Moschonas, I. F., Kappos, A. J., Panetsos, P., Papadopoulos, V., Makarios, T. and Thanopoulos,
P. (2008). Seismic fragility curves for greek bridges: methodology and case studies.
Bulletin of Earthquake Engineering. 7(2), pp:439-468.
Melek, M. and Wallace, J. W. (2004). Cyclic behavior of columns with short lap splices. ACI
Structural Journal, 101.
Menegotto, M. and Pinto, P. (1973). Method of Analysis for Cyclically Loaded RC Frames
Including Changes in Geometry and Non-elastic Behaviour of Elements Under
Combined Normal Force and Bending. IABSE Congress Reports of the Working
Commission.
Miller Jr, R. G. (1997). Beyond ANOVA: basics of applied statistics. CRC Press.
Moschonas, I. F., Kappos, A. J., Panetsos, P., Papadopoulos, V., Makarios, T., and P.
Thanopoulos (2008). Seismic fragility curves for greek bridges: methodology and case
studies. Bulletin of Earthquake Engineering 7(2), pp:439-468.
MTD (2008). Bridge Memo to Designers (10-20), California Department of Transportation,
Sacramento, CA.
234
Muthukumar, S. and DesRoches, R. (2006). A Hertz contact model with non-linear damping for
pounding simulation. Earthquake Engineering and Structural Dynamics, 35, pp:811-828.
Ni, P., Mangalathu, S., Mei, G., Zhao, Y. (2017) Permeable Piles: An Alternative to Improve the
Performance of Driven Piles, Computers and Geotechnics, Vol. 84, 2017.
Nielson, B. G. (2005). Analytical fragility curves for higway bridges in moderate seismic zones.
PhD. Thesis, Georgia Tech, Atlanta.
Nielson, B. G, and DesRoches, R. (2006). Influence of modeling assumptions on the seismic
response of multi-span simply supported steel girder bridges in moderate seismic zones.
Engineering Structures, 28, pp:1083-1092.
Ohtaki, T., Benzoni, G., Priestley, M., and Seible, F. (1997). Seismic performance of a full scale
bridge column-as built and as repaired. Second National Seismic Conference on Bridges
and Highways.
Orozco, G. L. (2001). The effects of a large velocity pulse on reinforced concrete bridge columns.
Dept. of Structural Engineering, University of California, San Diego.
Owen, S. V. and Froman, R. D. (1998). Focus on qualitative methods uses and abuses of the
analysis of covariance. Research in Nursing and Health, 21(6), pp:557–562.
Park, J. and Towashiraporn, P. (2014). Rapid seismic damage assessment of railway bridges
using the response-surface statistical model. Structural Safety, 47, pp:1-12.
Padgett, J. E. (2007). Seismic vulnerability assessment of retrofitted bridges using probabilistic
methods, Ph.D Thesis, Georgia Tech, Atlanta, Georgia.
Padgett, J. E., and DesRoches, R. (2007). Sensitivity of seismic response and fragility to
parameter uncertainty. Journal of Structural Engineering, 133, pp:1710-1718.
Porter, K. (2010). Cracking an open safe: uncertainty in HAZUS-based seismic vulnerability
functions, Earthquake Spectra, 26(3), pp:893-900.
Porter, K. A. (2003). An overview of PEER’s performance-based earthquake engineering
methodology. Proceedings of Ninth International Conference on Applications of
Statistics and Probability in Civil Engineering.
Potyondy, J. G. (1961). Skin friction between various soils and construction materials.
Geotechnique, 11(4), pp: 339-353.
Priestley, M. N., Seible, F., and Calvi, G. (1996). Seismic design and retrofit of bridges. John
Wiley and Sons, New York.
Priestley, M. N. and Benzoni, G. (1996). Seismic performance of circular columns with low
longitudinal reinforcement ratios. ACI Structural Journal 93(4).
Ramanathan, K. N. (2012). Next generation seismic fragility curves for California bridges
incorporating the evolution in seismic design philosophy. Ph.D Thesis, Georgia Tech,
Atlanta, Georgia.
Ramanathan, K., Jeon, J.-S., Zakeri, B., DesRoches, R., and Padgett, J. E. (2015). Seismic
response prediction and modeling considerations for curved and skewed concrete box-
girder bridges. Earthquakes and Structures, 9(6), pp:1153-1179.
Ramanathan, K., Padgett, J. E., and DesRoches, R. (2015). Temporal evolution of seismic
fragility curves for concrete box-girder bridges in California, Engineering Structures,
97, pp:29-46.
235
Ranf, R. T., Eberhard, M. O., and Malone, S. (2007). Post-earthquake Prioritization of Bridge
Inspections. Earthquake Spectra 23(1), pp:131-146.
Ranf, R. T., Nelson, J. M., Price, Z., Eberhard, M. O., and Stanton, J. F. (2006). Damage
accumulation in lightly confined reinforced concrete bridge columns, Pacific Earthquake
Engineering Research Center.
Roblee, C. J. (2016a). Memorandum: Box Girder bridges – span length models. P266-T1780
Technical Archives File Bridge Component Models, California Department of
Transportation, Sacramento, California.
Roblee, C. J. (2016b). Memorandum: Box Girder bridges – transverse profile models. P266-
T1780 Technical Archives File Bridge Component Models, California Department of
Transportation, Sacramento, California.
Saiidi, M., Maragakis, E., and Feng, S. (1996). Parameters in bridge restrainer design for seismic
retrofit. Journal of Structural Engineering, 122(1), pp:61-68.
Saini, A. and Saiidi, M. S. (2014). Probabilistic Damage Control Approach for Seismic Design
of Bridges. University of Nevada, Reno.
Sanchez, A. V., Priestley, M., and Seible, F. (1997). Seismic performance of flared bridge
columns. UC San Diago test report.
Schoettler, M., Restrepo, J., Guerrini, G., Duck, D., and Carrea, F. (2012). A full-scale, single-
column bridge bent tested by shake-table excitation. Las Vegas, NV: Center for Civil
Engineering Earthquake Research, Department of Civil Engineering, University of
Nevada.
Seo, J., and Linzell, D. G. (2012) Use of response surface metamodels to generate system level
fragilities for existing curved steel bridges. Engineering Strutures, 52, pp:642-653.
Shamsabadi, A. and Yan, L. (2008). Closed-form force-displacement backbone curves for bridge
abutment-backfill systems. Geotechnical Earthquake Engineering and Soil Dynamics IV.
2008. 1-10.
Shanmugam, S. P. (2009). Seismic behavior of circular reinforced concrete bridge columns
under combined loading including torsion, Ph.D thesis, Missouri University of Science
and Technology.
Shinozuka, M., Feng, M. Q., Lee, J., and Naganuma, T. (2000). Statistical analysis of fragility
curves. Journal of engineering mechanics, 126(12), pp:1224-1231.
Soleimani (2017). Fragility of California bridges – development of modification factors. Ph.D
Thesis, Georgia Tech, Atlanta, Georgia.
Soleimani, F., Mangalathu, S., and DesRoches, R. (2017) A comparative analytical study on the
fragility assessment of bridges with various column shapes, in Review.
Stefanidou, S. P., and Kappos, A. J. (2017). Methodology for the development of bridge-specific
fragility curves. Earthquake Engineering and Structural Dynamics, 46(1), 73-93.
Stone, W. C. and Cheok, G. S. (1989). Inelastic behavior of full-scale bridge columns subjected
to cyclic loading. Gaithersburg, MD, National Institute of Science and Technology.
Sun, Z., Priestley, M., and Seible, F. (1993). Diagnostics and retrofit of rectangular bridge
columns for seismic loads, Department of Applied Mechanics and Engineering Sciences,
University of California, San Diego.
236
Tibshirani, R, (1996). Regression shrinkage and selection via the lasso. Journal of Royal
Statistical Society 58(1), pp:267-288.
Vickers, A. J. (2005). Parametric versus non-parametric statistics in the analysis of randomized
trials with non-normally distributed data. BMC medical research methodology, 5(1), pp:
35.
Vidakovic, B. (2011). Statistics for bioengineering sciences: with MATLAB and WinBUGS
Support, Springer Science and Business Media.
Wald, D., Lin, K.-W., Porter, K. and Turner, L. (2008). ShakeCast: Automating and Improving
the Use of ShakeMap for Post-Earthquake Decision-Making and Response. Earthquake
Spectra 24(2), pp:533-553.
Yu, O., Allen, D. L., and Drnevich, V. P. (1991). Seismic vulnerability assessment of bridges on
earthquake priority routes in Western Kentucky. Lifeline Earthquake Engineering, pp.
817-826.
Zou, H. and Hastie, T. (2005). Regularization and variable selection via the elastic net. Journal
of Royal Statistical Socirty B 67, pp:301-320.
Zhang, J. and Huo, Y. L. (2009) Evaluating effectiveness and optimum design of isolation devices
for highway bridges using the fragility function method. Engineering Structures, 31,
pp:1648-60.
Zhong, J., Gardoni, P., Rosowsky, D., and Haukaas, T. (2008). Probabilistic seismic demand
models and fragility estimates for reinforced concrete bridges with two-column bents.
Journal of engineering mechanics, 134(6), pp:495-504.
Zelaschi, C., Monteiro, R. and Pinho, R. (2016). Parametric Characterization of RC Bridges
for Seismic Assessment Purposes. Structures, 7, pp: 14-24.
237
VITA
SUJITH MANGALATHU
PUBLICATIONS
Refereed Journal Articles
1. Ni, P., Mangalathu, S., Mei, G., Zhao, Y., “Permeable Piles: An alternative to
improve the performance of driven Piles”, Computers and Geotechnics, Vol. 84,
2017.
2. Mangalathu, S., Jeon J-S., DesRoches, R., Padgett, J. E., “ANCOVA-based grouping
of bridge classes for seismic fragility assessment”, Engineering Structures, Vol.
123, 2016.
3. Beyer, K., Mangalathu, S., “Numerical study on the peak strength of masonry
spandrels with arches”, Journal of Earthquake Engineering, Vol. 18, No. 2, 2014.
238
4. Beyer, K., Mangalathu, S., “Review of strength models for masonry spandrels”,
Bulletin of Earthquake Engineering, Vol. 11, No. 2, 2013.
5. Sujith, M. S., Menon, D., Dodagowdar, G. R., “Reliability analysis and design of
cantilever RC retaining walls against sliding Failure”, International Journal of
Geotechnical Engineering, vol. 5, No. 2, 2011.
8. Dukes, J., Mangalathu, S., Padgett, J. E., DesRoches, R., " Development of
bridge specific fragility methodology to improve the seismic resilience of
bridges", Earthquakes and Structures.
9. Jeon J-S., Mangalathu, S., Song, J., DesRoches, R., “Parameterized seismic
fragility curves for curved multi-frame concrete box-girder bridges using
Bayesian parameter estimation”, Journal of Earthquake Engineering.
10. Mangalathu, S., Jeon J-S., DesRoches, R., “Identification of Critical Parameters
on the Seismic Performance of Concrete Bridges using Lasso Regression”,
Earthquake Engineering and Structural Dynamics.
11. Ni, P., Mangalathu, S., Mei, G., Zhao, Y., “Compressive and flexural behavior of
Reinforced Concrete Permeable Piles”, Engineering Structures.
12. Nishanth, M., Dhir, P., Davis, R. Mangalathu, S., “Stochastic response of RC
buildings under high dimensional model representation”, Earthquake
Engineering and Structural Dynamics.
1. Mangalathu, S., Soleimani, F., DesRoches, R., Padgett, J. E., “ANOVA based
grouping of bridge classes”, 16th World Conference on Earthquake Engineering,
Santiago, Chile, 2017.
239
2. Mangalathu, S., Jeon J-S., Soleimani, F., DesRoches, R., Padgett, J. E., Jiang, J.,
“Sensitivity of fragility curves to parameter uncertainty using Lasso regression”,
16th World Conference on Earthquake Engineering, Santiago, Chile, 2017.
3. Mangalathu, S., Jeon J-S., DesRoches, R., Padgett, J. E., “Application of Bayesian
methods to probabilistic demand analyses of concrete box-girder bridges”,
Geotechnical and Structural Engineering Congress, Phoenix, Arizona, 2016.
4. Mangalathu, S., Jeon J-S., DesRoches, R., Padgett, J. E., “Analysis of covariance to
capture the importance of bridge attributes on the probabilistic seismic demand
model”, 10th Pacific Conference on Earthquake Engineering, Sydney, Australia,
2015.
5. Mangalathu, S., Jeon J-S., Soleimani, F., DesRoches, R., Padgett, J. E., Jiang, J.,
“Seismic vulnerability of multi-Span bridges: an analytical perspective”, 10th
Pacific Conference on Earthquake Engineering, Sydney, Australia, 2015.
6. Mangalathu, S., Jeon J-S., Soleimani, F., DesRoches, R., Padgett, J. E., “Fragility
analysis of box-girder bridges with lap-splice mode of failure”, EERI Annual
meeting, Boston, 2015.
2015 EERI travel grant for 2015 EERI Annual meeting, Boston.
MEMBERSHIPS
240
PROFESSIONAL ACTIVITIES
241