You are on page 1of 273

Time-dependent Resilience of Bridges under Multi-Hazard Scenarios

Submitted in partial fulfilment of the requirements of the degree of

Doctor of Philosophy

by

D. DINESH KUMAR

(174046002)

Supervisor:

Prof. Swagata Basu (née Banerjee)

Department of Civil Engineering

INDIAN INSTITUTE OF TECHNOLOGY BOMBAY

(2023)
Approval Sheet

This thesis entitled “Time-dependent Resilience of Bridges under Multi-Hazard


Scenarios” by D. DINESH KUMAR (174046002) is approved for the degree of Doctor of
Philosophy.

Examiners

____________________________

____________________________

____________________________

Supervisor

____________________________

Chairman

____________________________

Date: ________________

Place: ________________
Declaration

I confirm that this dissertation represents my own ideas and written phrases. If I have
included words or ideas from other sources, I have given proper credit by citing and referencing
the original sources. I also declare that I have adhered to all principles of academic ethics and
integrity and have not altered or fabricated any idea or information in this dissertation. I
understand that if I violate these conditions, the Institute may take relevant action against me.
The original sources may take legal action, if I haven’t cited them or obtained permission when
necessary.

____________________________
(Signature)

____________________________
(Name of the student)

____________________________
(Roll No.)

Date: ________________
ABSTRACT

Assessment of multi-hazard resilience for engineered systems, particularly bridges, has


gained significant attention in recent years. This dissertation focuses on this issue and explores
various aspects of time-variant resilience assessment of reinforced concrete (RC) bridges under
multi-hazard conditions. A comprehensive review of existing literature highlights key features
of this study, including the ways how multiple hazards interact to form multi-hazard conditions
for bridges, performance evaluation of bridges under multi-hazards, methods for loss
assessment, and approaches for post-event recovery. As observed, earthquake in the presence
of flood-induced scour and earthquake, corrosion and traffic are two very significant multi-
hazard conditions for inland bridges. In relation to the former multi-hazard scenario,
intensification of flood hazard characteristics attributed by regional climate change is studied
to investigate the additional threats to the safety and serviceability of river-crossing bridges.
To demonstrate, an existing bridge over the San Joaquin river in California is considered. Time-
variant risk and resilience of the bridge are evaluated, with future flood projections in the San
Joaquin river that is obtained from simulations using general circulation models and
macroscale hydrological models for this region. Projected design floods result in increased
scour around bridge piers, leading to a rise in risk and a drop in resilience compared to no-
climate-change scenario. To mitigate these consequences, ripraps are applied as a climate
change adaptation measure, and their size is determined based on the expected design flood
flow over the projection period. A cost-benefit analysis shows that this adaptive measure is
cost-effective throughout the remaining service life of the bridge.

For the same multi-hazard scenario, the study also investigates the role of key design
parameters related to piers of RC river-crossing bridges. These parameters include aspect and
longitudinal reinforcement ratio of RC piers and differential ground elevation between multiple
pier bents on water. Multi-hazard condition at the southeast part of Nepal is considered as
hazard condition at testbed. Regional seismic hazard is represented with a suite of earthquakes
generated based on regional seismic design spectra. Due to the regional flood hazard, expected
pier scour of investigated bridges are estimated from 100-year flood discharge (including
climate change projection) in the Koshi river, Nepal. Three-dimensional finite-element models
of chosen bridges, without and with scour, are developed incorporating scour and ±10%
variations in the design parameters. Fragility and risk curves of investigated bridges are
compared to assess the relative influence of the design parameters on bridge performance. The

i
findings highlight the significant influence of the aspect and longitudinal reinforcement ratio
of piers on the multi-hazard performance of riverine bridges. Research outcome also
demonstrates how design parameters can be revised to achieve risk-targeted multi-hazard
design of bridges.

Life-cycle performance and resilience of RC bridges in high-traffic, corrosive, and


seismically active regions may get significantly influenced by the multi-hazard condition of
earthquake, corrosion and traffic-induced fatigue. The coupled interaction of corrosion and
fatigue, referred to as corrosion–fatigue, causes gradual material degradation of bridge girders
and leads to an enhanced seismic vulnerability of bridges as deterioration proceeds. In this
context, the study here develops a numerical framework to estimate seismic resilience of two
RC bridges, having two and three spans, subjected to combined corrosion–fatigue degradation
over their lifecycles. Modelling of corrosion–fatigue involves the simulation of fatigue loads
on bridge girders using stochastic samples of gross vehicle weights obtained from weigh-in-
motion measurements and estimation of fatigue stresses at fatigue-critical locations of bridge
girders. Observed time-variant deterioration of bridge girders due to corrosion–fatigue is
utilized to develop finite-element models of these bridges, in which piers are also assumed to
have corrosion. Analyses result in an estimation of seismic vulnerability and resilience of these
bridges in the life-cycle context. Among the two bridges, a three-span bridge is observed to
experience higher reduction in resilience with time due to its higher flexibility than does a two-
span bridge. Overall, research outcome demonstrates the confronting role of corrosion–fatigue
degradation and emphasized its adverse impact on life-cycle seismic resilience of RC bridges.

Corrosion-fatigue degradation also results in a decrease in flexural stiffness of girders.


As a result, exposed bridges become weak in resisting earthquake loads, particularly for those
with significant vertical ground motion (VGM) components. Evaluating the long-term
performance of such bridges is crucial for seismic safety. This challenging scenario is studied
here using a representative multi-span RC bridge located in a region characterized by active-
faults. Here, the corrosion-fatigue degradation mechanism is modelled as a coupled
phenomenon of pitting corrosion of rebars and traffic-induced fatigue stresses that lead to
fatigue cracks at deep pit locations in rebars. A comprehensive numerical model of the bridge
incorporates all the essential features to accurately capture the gradual degradation throughout
its lifespan, as well as fluctuations in the response of key vulnerable bridge components under
horizontal and vertical excitations. Obtained results highlight that neglecting the impact of

ii
corrosion-fatigue degradation and VGMs can significantly increase the seismic vulnerability
of bridges, jeopardizing their safety during future earthquakes.

iii
ACKNOWLEDGEMENT

I would like to express my sincere gratitude to my dissertation advisor Prof. Swagata


Basu for giving me the admission to perform scientific research in this premium institute, which
I consider as remarkable opportunity to explore and understand the comprehension of civil
engineering (specifically bridge engineering). Since the beginning, she consistently gave her
support, exhibited patience and given encouragement throughout the completion of this
dissertation. I aspired her proficiency in both spoken and written English, especially when it
comes to prepare technical articles.

I would like to express my thanks to faculty members, Prof. Souvik Banerjee and Prof.
Jayadipta Ghosh, who were the members of my Research Progress Committee. Throughout
this journey, their constructive feedback provided during each annual progress seminar has
fuelled our research progress in completing this dissertation.

I would like to express my heartfelt appreciation and gratitude to the Science and
Engineering Research Board (SERB); Coalition for Disaster Resilience Infrastructure (CDRI),
and Industrial Research and Consultancy Centre (IRCC) at IIT Bombay for their generous
financial support, which has enabled me to pursue this academic journey from 2017 to 2023.

I would like to thank my lab mates, Dr. Sharanbaswa, Mr. Raj Arora, Mr. Suneet, and
Mr. Aditya, for making my time in Room no 115 of the Civil Engineering Department
enjoyable and for keeping me company. Their presence helped me feel less lonely. I also want
to express my gratitude to Mr. Sairam for helping me with my coursework and providing
technical insights in building numerical models for some of my research tasks.

Finally, I would like to express my gratitude to my family for their love, care, and
unwavering support in everything I do. They have my heartfelt love and appreciation, and also
thank everyone who contributed to my successful completion of my studies at IIT Bombay.

iv
Dedicated to all those who lead as an example

v
TABLE OF CONTENTS

ABSTRACT .............................................................................................................................. I

ACKNOWLEDGEMENT .................................................................................................... IV

TABLE OF CONTENTS ..................................................................................................... VI

LIST OF FIGURES ............................................................................................................. XII

LIST OF TABLES ............................................................................................................ XVII

LIST OF SYMBOLS ......................................................................................................... XIX

CHAPTER 1 INTRODUCTION ............................................................................................ 1

1.1 Background and Motivation ........................................................................................ 1

1.2 Objective and scope of research .................................................................................. 4

1.3 Research methodology ................................................................................................ 6

1.4 Organization of thesis.................................................................................................. 7

CHAPTER 2 EXISTING RESEARCH ON MULTI-HAZARD VULNERABILITY,


LOSS AND RESILIENCE ASSESSMENT OF HIGHWAY BRIDGES ......................... 10

2.1 Interaction of natural hazards .................................................................................... 10

2.1.1 Multi-hazard performance of bridges under the interaction of natural hazards. 11

2.1.2 Multi-hazard performance of bridges under the interaction of natural and


anthropogenic hazard ......................................................................................... 13

2.2 Highway bridges across river and in urban areas: Identifying the vulnerability ...... 14

2.2.1 Vulnerability model of bridges under Level-I interaction: Earthquake and flood-
induced scour ..................................................................................................... 15

2.2.2 Vulnerability model of bridges under Level-II interaction: Earthquake and


corrosion ............................................................................................................ 17

2.3 Definition of limit-states in the literature to develop fragility curves ....................... 19

2.4 Performance-based design criteria for highway bridges ........................................... 20

vi
2.4.1 Basics of performance-based seismic design framework .................................. 27

2.4.2 Existing research on multi-hazard risk assessment of bridges in the context of


PBD .................................................................................................................... 27

2.5 Life-cycle deterioration of embedded rebar in girder under the influence of corrosion-
fatigue .......................................................................................................................... 29

2.5.1 Existing fatigue life prediction model of corroded steel rebar .......................... 30

2.5.2 Existing the impact of corrosion-fatigue in RC structures: Experimental


approaches.......................................................................................................... 32

2.5.3 Existing the impact of corrosion-fatigue in RC structures: Analytical


approaches.......................................................................................................... 34

2.6 Vertical ground motion and corrosion-fatigue bridges: An essential connection ..... 36

2.6.1 Existing studies for investigating the effect of VGM on performance of


highway bridges ................................................................................................. 37

2.7 Seismic resilience of highway bridges ...................................................................... 39

2.7.1 Estimation of seismic resilience through quantitative approaches .................... 40

2.7.2 Estimation of seismic resilience through qualitative approaches ...................... 42

2.8 Post-event losses and recovery model for quantifying resilience ............................. 42

2.8.1 Existing direct loss models ................................................................................ 42

2.8.2 Existing indirect loss models ............................................................................. 44

2.8.3 Recovery model for assessing resilience ........................................................... 46

2.9 Literature gap ............................................................................................................ 49

2.10 Closure ...................................................................................................................... 50

CHAPTER 3 MULTI-HAZARD PERFORMANCE OF RIVER-CROSSING BRIDGES


IN CHANGING CLIMATE: RISK, RESILIENCE AND ADAPTATION ..................... 52

3.1 Integrated approach for the adaptation of bridges to changing climate .................... 52

3.2 Short description of case-study bridge ...................................................................... 54

3.3 Finite element model of the case-study river crossing bridge ................................... 56

vii
3.3.1 Prestressed girders ............................................................................................. 56

3.3.2 Piers (wall-type) ................................................................................................. 56

3.3.3 Pile foundations ................................................................................................. 57

3.3.4 Abutments .......................................................................................................... 58

3.3.5 Elastomeric bearing at in-span hinge ................................................................. 59

3.4 Regional flood hazard and its impact under changing climate (observed & projected)
................................................................................................................................... 59

3.4.1 Flood hazard curve (observed and projected) .................................................... 60

3.4.2 Flood-induced scour at bridge foundation ......................................................... 62

3.5 Regional seismic hazard ............................................................................................ 62

3.6 Change in ductility demand of pier under seismic demand ...................................... 64

3.7 Seismic vulnerability of bridge under changing climate ........................................... 66

3.7.1 Component-level fragility curve ........................................................................ 66

3.7.2 System-level fragility curve ............................................................................... 67

3.8 Multi-hazard risk and resilience of case-study bridge .............................................. 69

3.8.1 Direct and indirect losses ................................................................................... 69

3.8.2 Post-event recovery and restoration ................................................................... 70

3.9 Multi-hazard risk of case-study bridge...................................................................... 70

3.10 Multi-hazard resilience of case-study bridge ............................................................ 73

3.11 Discussion of results – impact of climate change on risk and resilience of the bridge .
................................................................................................................................... 74

3.12 Climate change adaptation ........................................................................................ 75

3.12.1 Application of rip-rap around bridge pier .......................................................... 76

3.12.2 Risk-based cost benefit of retrofitting................................................................ 77

3.13 Closure ...................................................................................................................... 81

CHAPTER 4 IMPACT OF BRIDGE DESIGN PARAMETERS ON MULTI-HAZARD


PERFORMANCE-BASED DESIGN OF RIVER CROSSING BRIDGES ...................... 84

viii
4.1 Multi-hazard condition at test bed region ................................................................. 84

4.1.1 Regional seismicity and selection of ground motions ....................................... 85

4.1.2 Characteristics of 100-year flood with climate change projection and scour
depth ................................................................................................................... 86

4.2 Description of representative bridge and design parameters .................................... 86

4.3 Impact of bridge design parameter on moment-curvature relation of pier ............... 88

4.4 Nonlinear time-history analyses of bridges and seismic fragility curves.................. 91

4.4.1 Limit state of various bridge component ........................................................... 91

4.5 Variation of component response under a given seismic demand ............................ 92

4.6 Seismic vulnerability of investigated bridges ........................................................... 93

4.6.1 Role of design parameters on bridge fragility curves ........................................ 95

4.6.2 Role of design parameters on multi-hazard risk of investigated bridges ........... 97

4.7 Closure .................................................................................................................... 102

CHAPTER 5 IMPACT OF COMBINED CORROSION-FATIGUE ON SEISMIC


VULNERABILITY AND RESILIENCE OF HIGHWAY BRIDGES ........................... 104

5.1 Overview of numerical framework for estimating seismic resilience of bridges


incorporating corrosion-fatigue degradation mechanism .................................................. 104

5.2 Modelling corrosion-fatigue degradation on RC girders ........................................ 105

5.2.1 RC bridge girder and traffic data ..................................................................... 105

5.2.2 Sampling of vehicles and stochastic GVWs .................................................... 107

5.2.3 Transient dynamic analysis considered vehicle-bridge coupled system.......... 110

5.2.4 Estimation of equivalent fatigue stresses ......................................................... 114

5.2.5 Effect of corrosion-fatigue on properties of steel rebar ................................... 117

5.3 Application to representative highway bridges ....................................................... 124

5.3.1 Representative bridges and finite element models........................................... 124

5.3.2 Nonlinear time history analyses of aging bridges ............................................ 127

ix
5.3.3 Impact of corrosion-fatigue on bridge seismic vulnerability and seismic
resilience of bridge ........................................................................................... 128

5.3.4 Impact of corrosion-fatigue on bridge seismic resilience ................................ 128

5.4 Closure .................................................................................................................... 130

CHAPTER 6 ANALYTICAL ASSESSMENT OF VERTICAL GROUND MOTION


EFFECT ON HIGHWAY BRIDGE UNDER THE IMPACT OF CORROSION AND
TRAFFIC INDUCED FATIGUE ....................................................................................... 133

6.1 Case-study region and a representative bridge ........................................................ 133

6.2 Intensity and vertical seismic component of 2001 Bhuj earthquake (Mw 7.7) ....... 136

6.3 Estimation of corrosion-fatigue damage on steel and concrete under site-specific


traffic load .................................................................................................................. 138

6.3.1 Observed truck traffic and its axle load statistics ............................................ 138

6.3.2 Corrosion-fatigue of rebar................................................................................ 141

6.3.3 Traffic-induced fatigue stresses ....................................................................... 144

6.3.4 Deterioration of rebar due to corrosion-fatigue accounting pitting and fatigue


cracks ............................................................................................................... 145

6.4 Finite element model to account the effect of VGM on bridge structure ............... 148

6.4.1 Elastomeric bearing ......................................................................................... 148

6.4.2 Pier ................................................................................................................... 149

6.4.3 Bridge deck and other components .................................................................. 150

6.5 Selection of ground motion for dynamic analysis (based on frequency content of 2001
Bhuj earthquake) ........................................................................................................ 151

6.6 Impact of corrosion-fatigue and VGM on life-cycle response of bridge ................ 153

6.6.1 Impact on natural frequency of bridge ............................................................. 153

6.6.2 Impact on seismic response of bridge girders .................................................. 154

6.6.3 Impact on seismic response of elastomeric bearing ......................................... 156

6.6.4 Impact on seismic response of bridge pier ....................................................... 159

x
6.7 Vulnerability analysis of corrosion-fatigued bridges under the combined impact of
VGM and HGM ......................................................................................................... 163

6.7.1 Selection of input ground motions to generate fragility curves ....................... 163

6.7.2 Intensity of selected ground motion................................................................. 164

6.7.3 EDP for the fragility analysis of bridge which are affected by corrosion-fatigue
and VGM ......................................................................................................... 168

6.7.4 Component level fragility curves of case-study bridge ................................... 168

6.7.5 System level fragility curves of case-study bridge .......................................... 172

6.8 Closure .................................................................................................................... 175

CHAPTER 7 SUMMARY AND CONCLUSION ............................................................ 177

7.1 Summary ................................................................................................................. 177

7.2 Conclusion............................................................................................................... 179

7.3 Key research contribution ....................................................................................... 182

7.4 Design recommendation for bridges in India .......................................................... 183

7.5 Limitation and future work ..................................................................................... 184

APPENDIX - A ..................................................................................................................... 186

APPENDIX - B ..................................................................................................................... 201

APPENDIX - C ..................................................................................................................... 203

PUBLICATIONS ................................................................................................................. 213

REFERENCES..................................................................................................................... 214

xi
LIST OF FIGURES

Figure 1.1 General flowchart to quantify risk and resilience of highway bridge ...................... 6
Figure 2.1 Level-I interaction (adopted and recreated based on Zaghi et al. 2016)………….12

Figure 2.2 Level-II interaction (adopted and recreated based on Zaghi et al. 2016) ............... 13
Figure 2.3 Flow-chart of PBD of bridges ................................................................................ 28
Figure 2.4 Different forms of expressing seismic risk of bridges............................................ 28
Figure 2.5 Dissolution of embedded rebar in girder due to corrosion-fatigue......................... 31
Figure 2.6 Ultimate strength for RC beams under static and cyclic loading ........................... 34
Figure 2.7 Conceptual illustration of seismic resilience .......................................................... 41
Figure 2.8 Direct loss estimation procedure using fragility curve ........................................... 44
Figure 2.9 Recovery models at different bridge damage states; (a) minor damage; (b) moderate
damage; (c) major damage; and (d) collapse state ................................................................... 48
Figure 3.1 General methodology to identify the adaptation solution for climate change……...53

Figure 3.2 Projected changes in environment and its impact on bridges due to climate change
.................................................................................................................................................. 54
Figure 3.3 Aerial view of case-study bridge (picture curtesy: Google Map) .......................... 55
Figure 3.4 Elevation view of case-study bridge....................................................................... 55
Figure 3.5 Schematic representation of 3D finite element model ........................................... 57
Figure 3.6 Back-bone curve of abutment in (a) longitudinal (b) transverse direction ............. 59
Figure 3.7 Annual peak discharge of the case-study river (observed and projected) .............. 61
Figure 3.8 Mean flood hazard curve for three flood scenarios; Flood Scenario I: Observed,
Flood Scenario II: Projected, and Flood Scenario III: Projected. ............................................ 61
Figure 3.9 (a) Seismic hazard curve of case-study region (b) PGA of selected ground motion
from PEER database ................................................................................................................ 63
Figure 3.10 Curvature ductility of a bridge pier under varied intensity seismic events for three
flood scenarios ......................................................................................................................... 65
Figure 3.11 Component-level fragility curves of case-study bridge for four damage states at (a)
Pier - Flood scenario I, (b) Pier - Flood scenario II, (c) Pier -– Flood scenario III, (d)
Elastomeric bearing at all flood scenarios, (e) Abutment active deformation at minor damage
state .......................................................................................................................................... 67
Figure 3.12 Seismic fragility curve of the bridge for three flood scenarios at (a) minor damage,
(b) moderate damage, (c) major damage and (d) collapse state .............................................. 68
xii
Figure 3.13 Expected losses for various combinations of the multihazard event .................... 72
Figure 3.14 Seismic risk curve of case study bridge for three scenarios of flood hazard ....... 73
Figure 3.15 Multihazard resilience of the bridge for the three scenarios of flood and seismic
hazard ....................................................................................................................................... 74
Figure 3.16 Elevation view of bridge pier retrofitted with rip-rap .......................................... 78
Figure 3.17 Section view of bridge pier retrofitted with rip-rap.............................................. 78
Figure 3.18 Benefit cost ratio (BCR) for rip-rap application around bridge piers as a climate
change adaptive measure; dark line represents mean BCRs estimated based on an average
discount rate of 2.2% and the shaded area is for mean±SD..................................................... 81
Figure 4.1 Generated earthquake ground motions based on the seismic hazard at the testbed
region…………………………………………………………………………………………86

Figure 4.2 Geometry, cross-sectional details, and major nonlinear components of the base
model........................................................................................................................................ 87
Figure 4.3 Moment-curvature relations of bridge piers for varied (a) diameter and (b)
longitudinal reinforcement ratio .............................................................................................. 91
Figure 4.4 Curvature ductility demand of bridge piers under seven strong ground motions;
bridges with (a) varied pier diameters, (b) varied longitudinal reinforcement ratio, and (c) same
or different riverbed elevation at pier bents. ............................................................................ 93
Figure 4.5 Seismic fragility curves of bridges without scour at (a) minor damage, (b) moderate
damage, (c) major damage, and (d) collapse state ................................................................... 94
Figure 4.6 Seismic fragility curves of bridges with scour at (a) minor damage, (b) moderate
damage, (c) major damage, and (d) collapse state ................................................................... 95
Figure 4.7 Percent variation in fragility median at various damage states due to (a) variations
in design parameters in non-scoured bridges, (b) variations in design parameters in scoured
bridges, and (c) combined effect of scour and variations in design parameter ........................ 97
Figure 4.8 Variation in multihazard risk with respect to that of the base model for bridges with
(a) varied pier diameter, (b) varied pier diameter in the presence of scour, (c) varied
longitudinal reinforcement ratio, and (d) varied longitudinal reinforcement ratio ................ 100
Figure 5.1 Flowchart for calculating material degradation of aging RC girders due to corrosion-
fatigue phenomenon………………………………………………………………………...105

Figure 5.2 (a) Dimension of the bridge deck at mid-span cross-section and (b) dimension of a
girder ...................................................................................................................................... 106

xiii
Figure 5.3 Statistics of GVW from different axle types; (a) Steering axle, (b) Single axle, and
(c) Tandem axle [Note: wi, µi, σi are parameters for ith axle representing, respectively, mixing
proportion, mean, and standard deviation of Gaussian Mixture Models, GMM].................. 109
Figure 5.4 Three-axle fatigue load model and its analytical representation for the ensuing
analysis ................................................................................................................................... 111
Figure 5.5 Results from the VBI analysis; (a) mid-span stress-time histories of fatigue-critical
rebars of the bridge girder under three GVWs and (b) comparison of bridge deflection obtained
from in-situ field test and the current study. .......................................................................... 115
Figure 5.6 Histogram of equivalent fatigue stress of rebar under varying GVW .................. 117
Figure 5.7 Projected growth in truck traffic volume based on the observed traffic .............. 118
Figure 5.8 Schematic representation of corrosion-fatigue deterioration of embedded rebar in
girder ...................................................................................................................................... 121
Figure 5.9 Time-dependent reduction of (a) residual area of rebar due to corrosion-fatigue and
(b) elastic modulus of concrete due to fatigue [Uncertainty bands (mean±1 SD) represent the
variability of mean estimates] ................................................................................................ 122
Figure 5.10 Time-dependent reduction of (a) moment of inertia of a T-girder and (b) flexural
stiffness of deck ..................................................................................................................... 123
Figure 5.11 Three-dimensional finite element model of the three-span bridge with linear and
nonlinear elements ................................................................................................................. 125
Figure 5.12 Comparison of load-deflection response of the bridge girder obtained from in-situ
test and the current study ....................................................................................................... 126
Figure 5.13 Seismic response of bearings for different traffic growth models ..................... 127
Figure 5.14 Time-dependent fragility curves of the three-span bridge at (a) minor damage, (b)
moderate damage, (c) major damage, and (d) collapse state (values within parenthesis represent
median values......................................................................................................................... 129
Figure 5.15 Life-cycle seismic resilience of the aging bridges; (a) two-span bridge and (b)
three-span bridge .................................................................................................................... 130
Figure 6.1 Tectonic features of Gujarat (Adopted from Chopra et al. [2012])………………134

Figure 6.2 Baseline dimension of representative T-girder bridge ......................................... 135


Figure 6.3 Acceleration time history in (a) N-S direction (b) E-W direction (c) U-D direction
................................................................................................................................................ 136
Figure 6.4 (a) Vertical-to-horizontal spectral ratio of 2001 Bhuj earthquake which recorded at
Ahmedabad station (b) Spectral acceleration of horizontal and vertical component of seismic

xiv
event (c) Fourier amplitude spectrum horizontal and vertical component of 2001 Bhuj
earthquake .............................................................................................................................. 137
Figure 6.5 Observed GVW of truck in a highway corridor of Gujarat .................................. 139
Figure 6.6 Statistics of observed axles in the sample – (a) Steering axle (b) Single axle (c)
Tandem axle ........................................................................................................................... 140
Figure 6.7 Conceptual representation of corrosion-fatigue damage mechanism of rebar ..... 142
Figure 6.8 (a) variation of equivalent stress with GVW (b) Gaussian mixture model of
equivalent stress ..................................................................................................................... 144
Figure 6.9 Demonstration (schematic) of area loss of rebar due to (a) uniform corrosion alone,
(b) the formation of deep pit (DP) and fatigue cracks after pit nucleation at pitting locations,
and (c) effective residual area of rebar under combined corrosion and fatigue..................... 146
Figure 6.10 Reductions in (a) residual area of steel rebar and (b) flexural stiffness of bridge
girder ...................................................................................................................................... 148
Figure 6.11 Representative bridge and its nonlinear elements .............................................. 152
Figure 6.12 Fundamental mode shapes and time periods of the bridge at pristine and 100-year
conditions ............................................................................................................................... 154
Figure 6.13 Response of a bridge girder under GM # 1 at the pristine and 100-year degraded
states; (a) fast Fourier transformation of girder acceleration in the vertical direction, (b) mid-
span vertical displacement, and (c) moment demand and estimated capacity ....................... 155
Figure 6.14 Maximum translation displacements of elastomeric bearings at the pristine and
100-year degraded states of the bridge; (a) bearing on exterior pier in a bent (b) bearing on
interior pier in a bent .............................................................................................................. 157
Figure 6.15 Variations in bearing response under GM # 1; (a) compressive stresses on an
exterior bearing, (b) compressive stresses on an interior bearing, and (c) coefficient of friction
of an interior bearing .............................................................................................................. 158
Figure 6.16 (a) Shear-deformation relationship of elastomeric bearing above the interior pier
and (b) variation of shear demand imposed on the underlying pier under GM # 1 ............... 159
Figure 6.17 Pier response at the pristine and 100-year degraded states; (a) yield curvature under
variable axial load ratios, (b) curvature ductility demands (CDDs) without the inclusion of
VGM component of GM # 1, and (c) curvature ductility demands (CDDs) with the inclusion
of VGM component of GM #1 .............................................................................................. 160
Figure 6.18 Curvature ductility demand of pier at pristine and 100-years of service under all
considered ground motions .................................................................................................... 162
Figure 6.19 Shear capacity of a bridge pier under six ground motions ................................. 162
xv
Figure 6.20 (a) Target and mean spectrum of selected ground motion from PEER NGA
database (b) V/H ratio of selected ground motions ............................................................... 164
Figure 6.21 Fragility curves for components such as (a) abutment (b) bearing (c) girder (d) pier
– in their pristine state and at different levels of damage. ..................................................... 170
Figure 6.22 Fragility curves for components such as (a) abutment (b) bearing (c) girder (d) pier
– in their 75th and 100 year of service life and at different levels of damage. ...................... 171
Figure 6.23 Time-dependent system level fragility curves at (a) minor (b) moderate (c) major
(d) collapse damage state ....................................................................................................... 173
Figure B.1 Cumulative distribution function (CDF) plot for annual peak discharge of the river
from observations (black) and (a) raw and (b) bias-corrected projections (other colours) during
the period of overlap between observations and model simulations (1950-2011) 202

xvi
LIST OF TABLES

Table 2.1 Limit state thresholds of cylindrical piers in terms of drift ..................................... 21
Table 2.2 Limit state thresholds of elastomeric bearings ........................................................ 23
Table 2.3 Limit state thresholds of abutments ......................................................................... 24
Table 2.4 Limit state thresholds of cylindrical piers in terms of curvature ductility ............... 26
Table 3.1 Calculated scour depth at bridge foundation……………………………………….62

Table 3.2 Damage state definition as given in HAZUS........................................................... 65


Table 3 3 Magnitude of Traffic Parameters Associated with Indirect Loss ............................ 71
Table 4.1 Specifications of the bridge models under consideration…………………………89

Table 4.2 Percentage change in the repair cost of non-scoured bridge models relative to
acceptable seismic risk under three design seismic events .................................................... 101
Table 4.3 Percentage change in the repair cost of scoured bridge models relative to acceptable
seismic risk under three design seismic events ...................................................................... 101
Table 5.1 Truck types and its percentage occupancy in observed traffic volume………….107

Table 5.2 Parameters associated with truck model for VBI analysis .................................... 112
Table 6.1 Observed peak ground acceleration (PGA) of aftershocks of 2001 Bhuj earthquake
(data obtained from Choudhury et al. 2016)…………………………………………………138

Table 6.2 Truck type and its percentage occupancy in the monitored samples ..................... 139
Table 6.3 Ground motions used for the nonlinear time history analyses of the case study bridge
................................................................................................................................................ 153
Table 6.4 Comparison of moment capacity and demand at the mid-span of bridge girder ... 156
Table 6.5 Non-structure specific IMs which are identified in literature ................................ 166
Table 6.6 Structure specific IMs which are identified in literature ....................................... 167
Table 6.7 Computed median fragility values of the bridge system and its associated components
for the 75th year service life .................................................................................................. 174
Table 6.8 Computed median fragility values of the bridge system and its associated components
for the 100th year service life ................................................................................................ 174
Table A.1 Synthesis of studies (up to 2018) on bridge performance assessment under multi-
hazard condition…………………………………………………………………………….186

xvii
Table C.1 Ground motions used to match seismic hazard spectrum in Nepal compatible with
10% exceedance in 50 years. ................................................................................................. 203
Table C.2 Ground motions used to match seismic hazard spectrum in Nepal compatible with
5% exceedance in 50 years. ................................................................................................... 205
Table C.3 Ground motions used to match seismic hazard spectrum in Nepal compatible with
2% exceedance in 50 years. ................................................................................................... 206
Table C.4 Recorded ground motions in Nepal obtained from the Center for Engineering Strong
Motion Data ........................................................................................................................... 208
Table C.5 Selected ground motions for seismic fragility analysis of RC highway bridge in the
state of Gujarat, India ............................................................................................................. 209
Table C.6 SRSS of vertical spectral acceleration of selected ground motions at different life-
cycle year of bridge ................................................................................................................ 211

xviii
LIST OF SYMBOLS

Symbol Definition

DSk kth damage state

RPFlood Recurrence interval of flood hazard

PGAj Peak ground acceleration of jth ground motion

Ck Median fragility parameter at kth damage state

rk Damage ratio of kth damage state

ζk Log-standard deviation of fragility curve at kth damage state

IM Intensity measure

∆𝑏𝑏 Drift ratio of pier which initiates rebar buckling

Lp Length of pier

∆𝑠𝑝𝑎𝑙𝑙 Drift ratio of pier which initiates spalling of concrete

hw Backwall height of abutment

pe Annual probability of seismic hazard

P Axial load

fck Characteristics concrete strength

Ag Gross sectional area of pier

D Diameter of circular pier

ke Coefficient for column shape

𝜌𝑙 Longitudinal reinforcement ratio of pier

𝜌𝑡 Transverse reinforcement ratio of pier

Nf Fatigue life of material

𝛥𝜎 Stress range

Exponent which is derived based on fatigue test of metal (Stress-


m
life equation)

xix
Symbol Definition

material constant which is derived based on fatigue test of metal


𝐴′
(Stress-life equation)
𝛥𝜀𝑝 Plastic strain range

Exponent which is derived based on fatigue test of metal (Strain-


q
life equation

material constant which is derived based on fatigue of metal


𝐵′
(strain-life equation)

R Seismic resilience

frec Recovery function

t0E Time of earthquake event

TLC Control time

PE (DS=k) Probability of damage at damage state k

Cop,car Operation cost of car

Cop,truck Operation cost of truck

ADT Average daily traffic

ADE Average daily truck remaining on the bridge

TRD Average daily truck traffic ratio

Dl Detour length

CTL Cost due to time loss

CAW Average wage per hour

Ocar Average car occupancy

Otruck Average truck occupancy

CATC Average total compensation per hour

Cgoods Time-value of goods transported in cargo

S Average detour speed

l Length of link

xx
Symbol Definition

SD Average speed of car on damaged Link

SO Average speed of car on intact Link

W Width of bridge deck

L Length of bridge

N Number of recovery options

CDL Risk due to physical damage

Creb Reconstruction cost

Crem Debris removal cost

Vc Coefficient of velocity that related to recovery

Proportionality factor between cost for construction of


bk
temporary bypass and bridge replacement

α, β, γ Coefficients related to outcomes of recovery action

Pabut Strength of abutment wall in longitudinal direction

Ae Effective area of abutment wall

𝐾𝑎𝑏𝑢𝑡,𝑙 Stiffness of abutment wall in longitudinal direction

𝐾𝑎𝑏𝑢𝑡,𝑡 Stiffness of abutment wall in transverse direction

Ki Initial stiffness of abutment

w Projected width of abutment diaphragm wall

hww Height of wing-wall of abutment

𝑃𝑎𝑏𝑢𝑡,𝑡 Strength of abutment wall in transverse direction

Cw Coefficient of participation factor

𝐾𝑠ℎ𝑒𝑎𝑟 Shear stiffness of elastomeric bearing

Fy Shear strength of elastomer

A Plan area of elastomer

G Shear modulus of elastomer

xxi
Symbol Definition

Tr Thickness of elastomer

𝜇 Friction coefficient between elastomer and surface of girder

𝑁𝑏𝑒𝑎𝑟𝑖𝑛𝑔 Axial force on elastomeric bearing due to gravity loading

y1 Upstream flow depth from the bottom of pier

a Width of pier

Correction factor for pier nose shape, angle of attack of flow and
K1, K2, K3
condition of bed
Fr1 Froude number of river flow

V Velocity of river flow

Ss Specific gravity of rip-rap

g Acceleration due to gravity

𝜇𝜑 Curvature ductility of pier

Δlong,a Longitudinal deformation of abutment in active-direction

Δlong,p Longitudinal deformation of abutment in passive-direction

Δb,long Deformation of elastomeric bearing in longitudinal direction

Ms Multi-hazard scenario

𝑑𝑟50 Median size of rip-rap

B Expected annual benefit

B Total benefit after retrofitting in remaining service life

Indirect losses of bridge under eth earthquake scenario (not


CSe0
retrofitted)

0
Direct losses of bridge under eth earthquake scenario (not
C Re
retrofitted)

CSeR Indirect losses of bridge under eth earthquake scenario (retrofitted)

R
CRe Direct losses of bridge under eth earthquake scenario (retrofitted)

xxii
Symbol Definition

Number of scenario earthquakes that represent regional


E
seismicity

DR Annual discount rate

T Remaining service life of the retrofitted bridge

CR Total cost involved in scour retrofitting

Mw Moment magnitude

M-ϕ Moment-curvature relationship

Mixing proportion of ith normal distribution in Gaussian Mixture


wi
Model

µi Mean of ith normal distribution in Gaussian Mixture Model

Standard deviation of ith normal distribution in Gaussian Mixture


σi
Model

 Rotational degree of freedom of fatigue truck

Vertical displacement of fatigue vehicle when traversing on


ys
bridge

yt Vertical displacement of tyre when traversing on bridge

Ks Stiffness of vehicle suspension

Cs Damping of vehicle suspension

Kt Stiffness of vehicle tyre

Ct Damping of vehicle tyre

MB Mass per unit length of bridge

CB Damping coefficient of bridge

I Moment of inertia of bridge

J Rolling moment of inertia of truck

E Young’s modulus of section

Mt Mass matrix of truck

xxiii
Symbol Definition

Kt Stiffness matrix of truck

Ct Damping matrix of truck

𝑦𝑡 Displacement vector of truck

𝑦̇ 𝑡 Velocity vector of truck

𝑦̈ 𝑡 Acceleration vector of truck

Ft Vector of wheel-bridge contact forces acting on the bridge

𝑦̇ 𝐵 First order derivative of bridge deflection

𝑦̈ 𝐵 Second order derivative of bridge deflection


....
yB Fourth order derivative of bridge deflection

Na Number of axles in fatigue load model

fi (t) Time-variant contact wheel force from ith axle on the bridge deck

 Dirac delta function

𝑊𝑠𝑖 Static gravity load of vehicle imposed on bridge deck

h(x) Vertical irregularity of road surface on bridge

Δ𝑛 Discretized spatial frequency

k Magnitude that defines the classification of road

𝑛0 Spatial frequency to model road irregularity

nmax Maximum theoretical sampling frequency

B Sampling interval

NH Number of terms in the simple harmonic function

𝜑𝑗 Random phase angle

𝑦𝐵 Mid-span deflection of bridge

𝜎 Tensile stress of embedded rebar in girder

zg Depth below neutral axis of T-girder

xxiv
Symbol Definition

∆𝜎𝐸𝑞 Equivalent stress

∆𝜎𝐷 Constant amplitude fatigue limit (CAFL)

KC and KD Coefficients to represent fatigue strength in S-N curve

Number of stress cycles that are higher than and lesser than
nq and nr
CAFL

∆𝜎𝑞 Stress range which is higher than or equal to CAFL

∆𝜎𝑟 Stress range which is lower than or equal to CAFL

Tini Time for corrosion initiation

𝑡𝑧 z years after corrosion initiation time

Xcov Cover depth

Ccr Critical chloride concentration

Cs Surface chloride concentration

Dc Coefficient of diffusion

Residual area of embedded rebar after z years from corrosion


A(tz)
initiation

dini Initial diameter of rebar

𝑟𝑐𝑜𝑟𝑟 Corrosion rate

 s ,max Maximum nominal stress on rebar

fy Yield strength of rebar

As Cross-sectional area of rebar when rebar initiates corrosion

ni Number of cumulative traffic cycles at ith year

 ic Change in rebar area due to corrosion at ith year

 if Change in rebar area due to corrosion-fatigue at ith year

s Degree of corrosion

xxv
Symbol Definition

𝜀𝑐𝑒 Elastic strain in concrete

𝜀𝑐𝑝 Plastic strain in concrete

𝜎𝑐,𝑟𝑚 Relative mean stress of concrete

𝜎𝑐,𝑚𝑎𝑥 , 𝜎𝑐,𝑚𝑖𝑛 Maximum and minimum stress in concrete

ttruck Time duration of truck passage on bridge (in hrs)

∆𝜎𝑐,𝑟 Relative stress range of concrete


𝑓
𝐸𝐶 Apparent change in modulus of concrete due to fatigue

Mw Moment magnitude

Correction coefficient related to curing and environmental


𝑘𝑐 , 𝑘𝑒
factors

𝐷𝑐𝑙,0 Reference diffusion coefficient

𝑛𝑐𝑙 Age exponent

Rpf Pitting factor

p(t) Pit depth at time t

a Crack length

𝛥𝐾 Alternating stress intensity factor

𝐶𝑝 , m Fatigue material constant

f Frequency of cyclic load

𝐷𝑓 Dimensionless factor for notched geometry

w(t) Width of corroded pit at time t

𝑝𝐶𝐹 (𝑡) Pit depth at time t due to corrosion-fatigue

ADP Residual area of rebar after undergoing pitting corrosion

𝐴𝑖 Pristine area of rebar before corrosion

𝐴𝑈 Residual area of rebar (uniform form of corrosion)

xxvi
Symbol Definition

𝜃1 ,𝜃2 Geometric variables to quantify residual area of rebar after pitting

Residual area of rebar after undergoing corrosion-fatigue


𝐴𝐶𝐹
deterioration

𝐾𝑟𝑜𝑡 Rotational stiffness of pier

Average tension between the surface of concrete and longitudinal


u
rebar

EIeff Effective flexural rigidity of pier

𝑇𝑚 Mean time period of ground motion

ci Coefficient of Fourier amplitude

coefficient of discrete fast Fourier transform frequency of


fi
accelerograms

𝑢𝑔 Ground displacement

𝑢̇ 𝑔 Ground velocity

𝑢̈ 𝑔 Ground acceleration

T Total time duration of ground motion

𝑛 Number of vibration mode of bridge

𝑆𝑎 Spectral acceleration

𝑆𝑣 Spectral velocity

𝑃𝑆𝑣 Pseudo spectral velocity

xxvii
CHAPTER 1

INTRODUCTION

1.1 Background and Motivation

Highway bridges play an important role in promoting economic growth and community
development by providing safe, economic, and efficient transportation. A large number of
highway bridges worldwide are designed and built using provisional standards that ignore the
multi-hazard concept in structural design. According to the National Disaster Management
Plan (NDMP 2019), 59% of Indian territory is susceptible to moderate and severe earthquakes,
and it is likely to result in disruption of life and properties in the form of physical and economic
consequences (or socio-economic vulnerability). The anticipated losses are expected to elevate
further by geoclimatic condition, environmental degradation, and urbanization. Sahana et al.
(2021) and Karuppusamy et al. (2021) have demonstrated the vulnerability of coastal villages
in India to multi-hazard. The study indicates that over 60% of the sampled villages in coastal
region of India are at severe risk from a combination of cyclones and tsunamis. Similarly, Roy
and Matsagar (2021) have listed few districts in India that are likely to face significant multi-
hazard risk arising from the interaction of natural hazards. Given these challenges, it is essential
to prioritize multi-hazard considerations in design and construction of bridges to ensure safety
and resilience in the face of multi-hazards.

Physical damage to bridges may cause temporary shutdown of transportation network,


and thus gives rise to socio-economic consequences. Damaged bridges require repair and
replacement, which can pose burden on expenditure and result in reduced economic growth.
Understanding the physical vulnerability of highway bridges to natural disasters is crucial for
engineers and policy-makers to take proactive measures to minimize the failure of bridge and
reduce its associated consequences through risk assessment. By reducing the physical
vulnerability of highway bridges to disaster event, transportation infrastructure can remain
functional and become more resilient in the face of disaster event. In general, understanding a
multi-hazard scenario for bridge infrastructure requires a primary hazard along with an
interacting hazard (also termed as secondary hazard) such that these together influence the

1
structural performance of bridges through complex interactions. For an example, bridges on
rivers are susceptible to multi-hazard scenario arising due to earthquake and flood. Flood
causes scouring and results in undermining of soil and sediments around bridge foundations.
This, in turn, compromises the flexibility of bridge, and results in increased vulnerability to
seismic hazard. Hence, the combined effect of an earthquake and a flood can result in complete
collapse. The occurrence of earthquake and flood are independent but its interaction impose
higher vulnerability to bridge structure. Furthermore, global climate change can increase the
intensity and frequency of flooding events, resulting in increased vulnerability of bridge
structures (Yang and Frangopol 2019). As a result, the combination of earthquake and flood
pose a significant challenge to practicing engineers and policy-makers to consider and prepare
for risk mitigation under the effect of multi-hazard.

Another general example of independent-interacting multi-hazard scenario is


earthquake-corrosion. Highway bridges which are located in aggressive environment are
subjected to environmental corrosion and its constituent materials are prone to rigorous
deterioration. This alters the condition of the bridge and results in reduced performance during
seismic events. Therefore, it is crucial to account for this mechanism to assess and manage
seismic risk effectively throughout the service life of bridge. The recent collapse of the
Surajbari bridge in the salty environment of Gujarat (India) in the 2001 Bhuj earthquake was
caused by deterioration of reinforcement (Madabhushi and Haigh 2005). Corrosion and traffic
can both also lead to reinforcement deterioration, which is commonly referred to as corrosion-
fatigue. As a result, earthquake-corrosion-traffic is also an important independent-interacting
hazard scenario that exacerbates each other’s effect, leading to significant damage and
increased risk of bridge failure. As a result, in order to ensure the social and economic well-
being of a community, it is a dire necessary to assess the multi-hazard vulnerability of existing
highway bridges that are under likely to exposed to two or more interacting hazards.

In recent times, the rise in extreme hydrological events such as floods, mean seal level
rise and hurricane due to climate change has attributed to increase in vulnerability of bridge
structures (Yang and Frangopol 2019). Increase in severity of hydrological events due to
climate change can results in higher scour depth. Scour is identified as a general cause of bridge
failures (Cook et al. 2015). There are several empirical mathematical relationships available in
the literature to quantify scour depth around piers. These relationships are related to flow
discharge, velocity, water height. Climate change has the tendency to amplify all of above
variables. As a result, the risk of scour will increase. Several studies have addressed the

2
influence of climate change on scour risk (Dikanshi et al. 2016; Kallias and Imam 2016);
however, studies to assess the impact of scour risk on seismic performance of river crossing
bridges is rather low. This literature gap needs to be addressed by investigating the potential
risk of climate change on seismic risk and resilience of highway bridges, and proper adaptation
measure(s) to reduce the seismic risk should be proposed. Bridges are designed and built to last
for an extended period of time (sometimes exceeding 75 years). As a result, there is an urgent
need to ascertain its reliable performance in its early design phase for climate-change risk.

Limit state design (LSD) was used to modernize Indian bridge designs. This design
improved the safety and serviceability of bridges to extreme loading events such as
earthquakes. Bridges built in the late twentieth century are based on LSD. The design ensures
that bridges will not collapse or deform significantly under extreme loads. However, the design
philosophy is gradually evolving such that bridges will remain functional during and after an
extreme event, rather than avoiding collapse. This holistic way of design approach which
ensures the desired level of performance or function under various extreme loading conditions
is known as performance-based design (PBD). There is a knowledge gap on investigating
the specific roles of bridge design parameters related to pier for multi-hazard design of river
bridges in the context of achieving the desired performance level ranging from 'fully
operational' to 'collapse'.

As discussed earlier, bridges are designed for 75 years of service life. Transport vehicles
such as light and multi-axled trucks have grown rapidly in developing countries over the last
three decade. According to the annual report of Ministry of Road and Highway Transport of
India (MoRTH 2022), the total cumulative number of registered vehicles (both transport and
non-transport) increased to 296 million in 2018-2019, accounting a compound annual growth
rate (CAGR) of 9.91%. The category of transport vehicles (light and multi-axled trucks)
accounts for 8.8 percent of total registered vehicles. Increased frequency of truck traffic can
result in fatigue condition to bridges due to repeated stress and strain caused by trucks which
are passing over the bridge. This scenario along with environmental corrosion can lead to
deterioration of rebar (or reinforcing steel) over time (Blikharskyy et al. 2021). Deterioration
of rebar in reinforced concrete (RC) girders can have a significant impact on the seismic
performance of the bridge. As the geometric area of rebar decreases, it loses strength and
stiffness of the section, resulting in a lower capacity to withstand extreme loads. This causes
increased deformation or even failure of the bridge during a seismic event. As a result, the
mechanism of corrosion and fatigue must be considered in the design phase of bridges to ensure

3
safety and reliability in the event of a multi-hazard scenario. There is, however, almost no
literature or existing development that takes into account the evolution of truck traffic growth
in the seismic vulnerability of highway bridges. As a result, a systematic assessment of multi-
hazard vulnerability that takes earthquake-corrosion-traffic into account should be proposed,
which will aid in the improvement of seismic risk associated with bridges for future hazard
events.

1.2 Objective and scope of research

The overarching objective of this research is to develop suitable analytical frameworks


to better understand the interactions of various multi-hazard conditions and their effects on
vulnerability, risk, and resilience of highway bridges. Two major multi-hazard conditions for
highway bridges – earthquake and flood-induced scour and earthquake, corrosion, and traffic
– are investigated in detail. Under the stated objective, specific tasks and scope of the research
are discussed below.

Task 1. Impact of climate change on multi-hazard performance of highway bridges: risk,


resilience and adaptation

• An integrated approach is presented to assess the impact of global climate change


on bridge performance under multi-hazard scenario involving floods and
earthquakes.

• Flood frequency analysis are performed for future flood events that can get
enhanced due to climate change.

• Multi-hazard performance of an existing highway bridge is evaluated in terms of


its risk and resilience.

• To reduce future monetary consequences, a climate change adaptation measure for


the bridge is proposed.

• Cost-benefit analysis is performed for the proposed adaptation measure over the
remaining service life of the bridge.

Task 2. Effect of bridge design parameters on multi-hazard performance of river crossing


bridges

• Impact of various pier design parameters on multi-hazard performance of highway


bridges involving earthquake and flood-induced scour is investigated.

4
• Impacts of investigated parameters are measured in terms of the change in bridge
fragility characteristics and multi-hazard risk.

• Further, bridge performance at desired level of performance ranging from ‘fully


operational’ to ‘collapse’ is studied.

Task 3. Influence of combined corrosion-fatigue deterioration on life cycle resilience of RC


bridges

• A quantitative framework to account for real life traffic data and environmental
hazard (i.e., corrosion) on life-cycle performance of bridges is proposed.

• Various analysis modules such as generation of stochastic traffic load for the study
region, vehicle-bridge interaction (VBI) to obtain dynamic stresses of bridge
girder, extracting fatigue stress from the obtained stress-time history, and material
deterioration owing to coupled corrosion-fatigue phenomenon are evaluated.

• A detailed finite element model (FEM) of bridge girder is developed to account for
corrosion-fatigue deterioration along bridge service life and its accuracy is
validated using information from field test data.

• The influence of corrosion-fatigue is represented on life-cycle seismic resilience of


the investigated bridge.

Task 4. Dynamic behaviour of RC bridges under corrosion-fatigue damage and near-fault


conditions

• An analytical formulation for mathematically modelling the residual area of rebar


due to pitting corrosion and fatigue cracks is proposed.

• Numerical simulations are performed to study the change in dynamic behaviour of


a RC T-girder bridge under the influence of combined deterioration owing to
corrosion-fatigue.

• Changes in dynamic behaviour of the bridge deck, elastomeric bearings, and piers
under near-fault ground motions of the representative bridge after 100 years of
service are investigated in comparison to that at its pristine condition.

5
Task 5. Influence of vertical ground motion and corrosion-fatigue damage on the seismic
vulnerability model of RC highway bridges

• Influence of vertical ground motion (VGM) on seismic vulnerability of various


component of a RC highway bridge with corrosion-fatigue degradation is
investigated for every 25-year interval throughout the bridge service life.

• System level vulnerability model of the bridge is developed. This enables the
identification of key bridge components that have significant impacts on overall
damage condition of the bridge.
Note that scope of the study is limited to estimating the local scour depth around the
foundation caused by one-time flood events. The study does not take into account of time-
dependent scour or other scenarios such as the impact of floating debris on the substructure and
its associated change in dynamic response of bridge. Only straight bridges, without skewness
and curvature, are analysed here for multi-hazard performance evaluation.

1.3 Research methodology

A general flowchart to assess multi-hazard vulnerability, risk, and resilience of highway


bridges is shown in Fig. 1.1. In this multi-step process, the methodology consists of various
modulus that require identifying and assessing the hazard, quantifying its potential
consequences on the community due to bridge damage, and expressing the same in the form of
risk curves. The potential of a damaged bridge to withstand and recover from the disastrous
event is quantified and expressed in the form of resilience.

Figure 1.1 General flowchart to quantify risk and resilience of highway bridge

6
1.4 Organization of thesis

The thesis is divided into seven chapters. Following this background information and
statement of problem, Chapter 2 discusses the overview of past research on multi-hazard
vulnerability, loss, and resilience assessment of highway bridges. Additionally, the chapter also
highlights the literature gap that will be addressed in this dissertation.

Chapter 3 discusses the multi-hazard performance of highway bridges under the


influence of earthquake and flood-induced scour that may get further enhanced due to climate
change. The impact of global climate change on enhancement of futuristic flood event is
presented as hazard curve. The chapter proposes an integrated method for assessing the impact
of climate change on the seismic performance of highway bridges due to the presence of local
scour around bridge foundations in the event of a future flood. The proposed approach is
demonstrated on existing highway bridge in California. The multi-hazard performance of
highway bridge is evaluated in terms of risk and resilience. To reduce the anticipated risk and
improve the resilience of highway bridges for futuristic flood events, adaptation measure is
proposed, and its monetary benefits for the remaining service life of the highway bridge is
analysed through cost-benefit analysis.

Chapter 4 investigates the influence of design parameters related to pier on multi-


hazard performance of highway bridge involving earthquake and flood. The investigated
design parameters are – aspect ratio and longitudinal reinforcement ratio of pier, and
differential ground elevation between multiple bent, and ± 10% variations of the stated design
parameters are used to investigate the variability on seismic performance of a highway bridge.
The investigation is carried out on a representative highway bridge, which is assumed to be
located in the southwest part of Nepal. Regional seismic hazard is represented by a set of
ground motion records which are gathered from various sources, and the regional flood hazard
of the Koshi River in Nepal is obtained from published literature that contains 100-year flood
discharge considering the effect of climate change. Relative influence of design parameters on
seismic performance of considered bridges is evaluated in terms of fragility and risk curves.
Finally, a performance-based approach is adopted to demonstrate the influence of design
parameters on risk-targeted multi-hazard design of bridges.

Chapter 5 discusses the multi-hazard performance of highway bridges under the


influence of earthquake-corrosion-traffic. Corrosion and traffic-induced fatigue are two major
threats to existing highway bridges. The coupled interaction of corrosion and fatigue causes

7
corrosion-fatigue condition and deteriorates the geometric and material property of embedded
rebar and concrete in RC girders. Reductions in these properties reduce the reserve strength of
existing bridges and increase its seismic vulnerability. This chapter proposes a numerical
framework to quantify the seismic vulnerability of RC bridges under the condition of corrosion-
fatigue over their life-cycle years. Two representative bridges with varying span number and
overall length are used to demonstrate the proposed framework. The quantified vulnerability is
used to estimate and compare life-cycle resilience of the case-study bridges for varying seismic
events.

Chapter 6 discusses the improved deterioration model of embedded rebar in RC girder


which is subjected to corrosion-fatigue degradation. The integrated process of pitting corrosion
and fatigue cracks on the residual area of rebar due to corrosion-fatigue is incorporated into the
analytical model. The damage is calculated using traffic and corrosion data from the Gujarat
region in India. The investigation focuses on how corrosion-fatigue degradation and VGM
jointly affect dynamic behaviour of the deck, elastomeric bearing, and pier of a typical
representative RC bridge in Gujarat, taking into account of deterioration in girder and pier. The
study also develops component-level fragility curve of the bridge for every 25-year interval.
Using component-level fragility curves, system-level fragility curves (i.e., vulnerability model)
are developed by identifying dominant bridge components that are responsible for various
damage state of the bridge throughout its service life.

Chapter 7 summarizes and concludes the research with key conclusions and
contributions, and discusses the direction for future research work.

8
This page is intentionally left blank

9
CHAPTER 2

EXISTING RESEARCH ON MULTI-HAZARD VULNERABILITY,


LOSS AND RESILIENCE ASSESSMENT OF HIGHWAY BRIDGES

Safety and serviceability of highway bridges in the face of multi-hazard events are
critical to ensure continued operation of highway transportation networks. The multi-hazard
scenario for bridge may arise from the interaction of natural and natural event, or natural and
non-natural event, and this interaction can alter the condition of a bridge and results in reduced
structural performance. Performance of a bridge structure under the influence of a natural event
is directly linked with structural strength, stiffness, ductility, and energy dissipation capacity
of various bridge components. This chapter discusses how natural events interact with each
other and impact bridge components, leading to reduced strength and stiffness, and impaired
performance during seismic activities. It also discusses an overview of possible multi-hazard
scenarios and highlights the state-of-art-research focused on the vulnerability evaluation of
bridges and methodologies to evaluate resilience of bridges in the aftermath of disaster event.

2.1 Interaction of natural hazards

The concept of multi-hazard scenario involves a primary hazard that can be interacted
by additional secondary hazard, and resulting in more severe destruction. Mahmoud and Cheng
(2017) classified the combination of multi-hazard into various group such as simultaneous and
correlated (e.g., earthquake and flood-induced-scour), simultaneous and uncorrelated (e.g.,
earthquake and snow), non-simultaneous and uncorrelated (e.g., earthquake and wind), and
correlated and cascading (e.g., fire following blast). Based on hazard interaction and its
mechanism and frequency of occurrence, Zaghi et al. (2016) classified multi-hazard condition
as level-I and level-II interactions. The level-I interaction involves hazards that naturally
interact with each other, either concurrently or successively. These interactions are referred to
as concurrent or successive hazards. Multiple hazards that occur at the same time or overlap
for a period of time are referred to as concurrent hazards. The duration of the overlap may have
a direct influence on structural performance of bridges. Typical examples of concurrent hazards
are storm surge, waves, and high wind that form and concurrently exist during hurricane (Ataei
and Padgett 2013). A successive hazard is one that triggers another, acts like a domino, and

10
results in the formation of multiple hazards. Typical example of this type is earthquake and
earthquake-trigged tsunami (Akiyama et al. 2013). The other possible interaction between
natural hazards is presented as design interaction matrix in Fig. 2.1. In this figure, a primary
hazard refers to a natural hazard that occurs independently, while a secondary hazard refers to
a natural hazard that occurs concurrently or successively with primary hazard.

Level II interaction between natural hazards occurs via their effects. To characterize the
interaction of hazards, Zaghi et al. (2016) classified four types of effects – site effects, physical
impacts, network and system disruptions, and social and economic consequences. The term site
effects define the hazard and represent the actions caused by the hazard at a specific location.
They must be understood in the context of physical impact. These effects are not dependent on
the presence of a bridge structure. Ground vibrations and liquefaction/ground settlement are
examples of site effects for earthquake. Physical impacts refer to the changes in the
behaviour of a physical component that are thought to be caused directly by one or more site
effects of natural hazards. For example, storm surges cause flooding, which in turn leads to
scouring around bridge foundations. As a result of scouring, the flexibility and dynamic
behaviour of bridges are subjected to change, thereby affecting their response to other natural
hazards. Disruptions to a system or network on a large scale are referred to as network and
system disruptions. For an example, flooding which inundates bridges may result in a
breakdown of transportation in network scale. Finally, the social and economic consequences
of hazards influence the human-societal behaviour. For an example, floods which inundate the
properties of local population can cause mass migration to other locations. The design
interaction matrix presented in Fig. 2.2 illustrates interaction between natural hazards, which
results in Level II interaction. The state-of-the-art research of multi-hazard performance
of highway bridges available in existing literatures is reviewed in two sections: (i) one that
forms and results from the interaction of natural hazards, and (ii) another that forms and results
from the interaction of natural and non-natural hazards (also referred as anthropogenic hazard).
The following section provides a brief summary of the review.

2.1.1 Multi-hazard performance of bridges under the interaction of natural hazards

Multiple number of studies are performed till date on highway bridges in the context of
exploring the structural behaviour of highway bridges under multi-hazard scenario arising from
the interaction presented in Fig 2.1 and Fig 2.2. Table A1 in Appendix A of this dissertation
presents a summary of these studies conducted in the context of multi-hazard with a list of its

11
intensity measures that are used to quantify the structural performance. The table comprises of
several combinations of hazards such as earthquake and flood-induced-scour, earthquake and
corrosion, earthquake and liquefaction, earthquake and hurricane, and among few others. These
combinations are primarily focused on modelling aspects related to developing fragility curves
and surfaces. Numerous approaches have been done in the literature in the context of analysis
and design of highway bridges to withstand multi-hazard, such as risk analysis (Kameshwar
and Padgett 2014), influence of aging on bridge performance (Ghosh and Padgett 2010),
structural stability of bridge pier under flood-induced scour (Chang et al. 2014), and strategies
for intervention (Deco and Frangopol 2011).

Figure 2.1 Level-I interaction (adopted and recreated based on Zaghi et al. 2016)

12
Figure 2.2 Level-II interaction (adopted and recreated based on Zaghi et al. 2016)

2.1.2 Multi-hazard performance of bridges under the interaction of natural and


anthropogenic hazard

In addition to exploration on natural multi-hazard scenarios on structural performance


of bridges, studies have also performed on the evaluation of bridges to explore its vulnerability
when exposed to multi-hazard event involving between natural and anthropogenic hazardous
event. Studies are not extensive as those listed in Table A1. However, the latter includes
combinations such as earthquake and blast (Bruneau et al. 2003), earthquake and traffic,
corrosion and traffic, scour and traffic, and barge-collision and scour. Studies have been
performed to investigate the vulnerability and risk of highway bridges under multi-hazard
scenarios which include simultaneous occurrence of traffic load, earthquake and flood-induced
scour. These investigations are performed in Ghosh et al. (2014), and Kameshwar and Padgett

13
(2018). Zhu and Frangopol (2013, 2016a, 2016b) conducted studies to evaluate failure
probability over time for different components of bridge superstructures, such as deck and
girders, in response to increased traffic flow.

In addition, the study also evaluated the failure probabilities of bridge substructures,
specifically the piers, in response to a hazard combination such as earthquake and scour. These
analyses were carried out in a multi-hazard scenario that accounted for the effect of traffic
loading. The multi-hazard risk of bridges is evaluated by considering the possibility of failure
of bridge components at both super and substructure levels, which may occur independently or
concurrently, leading to failure or closure of one or more lanes. The risk assessment involves
combining the risk linked to the failure of each individual bridge component. Previous research
has not established a clear trend in the seismic fragility of bridges under traffic load. However,
Ghosh et al. (2014) determined the deteriorating trend of seismic fragility with increasing truck
load intensity. Conversely, Kameshwar and Padgett (2018a) concluded that seismic fragility
can be either positively or negatively impacted by truck loading, depending on factors such as
bridge characteristics and traffic loads. Kameshwar and Padgett (2018b) also examined the
effect of multiple soil layers on bridge fragility under traffic and scour, while the other
(Kameshwar and Padgett 2018c) investigated the performance of scoured bridges due to barge
collision. Also, several studies were conducted over the past two decades on time-variant
failure probabilities and reliability of aging bridges under multi-hazard load cases involving
traffic and corrosion. This research includes studies by Hajializadeh et al. (2016), Stewart and
Rosowsky (1998), Val et al. (2000), and Vu and Stewart (2000). Till now, there has been a lack
of attention paid in literatures in assessing the seismic performance of RC bridges that undergo
deterioration caused by both environmental corrosion and traffic-induced mechanical stresses.
Furthermore, there is currently no established framework available in the literature to quantify
the change in vulnerability, seismic risk and resilience of these bridges under the multi-hazard
interaction of earthquake, corrosion and traffic over time.

2.2 Highway bridges across river and in urban areas: Identifying the vulnerability

Highway bridges that are built across rivers and located in urban areas are susceptible
to various kinds of interactions of natural hazards and external stressors, which can impact their
seismic vulnerability throughout their service life. Some of the common external stressors and
natural hazards that can affect their seismic vulnerability are identified and outlined below.

14
• Flood-induced scour: Bridges located across rivers are susceptible to floods that can
undermine soil around bridge foundations. This condition is known as scour and can
have a significant impact on seismic vulnerability of bridges.
• Environmental factors: Change in global temperature results in alternation in climate
and weather pattern, which can increase the likelihood of more frequent and intense
flood events in future. These flood events result in unanticipated levels of scour depths,
which can be higher than the design scour depth.
• Corrosion: Bridges in urban areas and in close proximity to chloride sources (such as
sea) are susceptible to corrosion, which may result in degradation of geometric area of
embedded steel in concrete, and weaken the structural integrity over time.
• Increased traffic flow and density: Bridges in urban areas are subjected to increased
traffic flow and results in fatigue over time. The repeated stress and strain from heavy
trucks results in fatigue cracks and deterioration of embedded rebar in RC girder.

These multi-hazard scenarios for river-crossing bridges in urban areas can occur
through either level-I interaction (i.e., earthquake and flood-induced scour) or level-II
interaction (i.e., earthquake, corrosion, and traffic). The following subsection discusses the
existing vulnerability models for multi-hazard scenarios involving earthquake, flood-induced
scour, corrosion and traffic.

2.2.1 Vulnerability model of bridges under Level-I interaction: Earthquake and flood-
induced scour

In general, the likelihood of concurrent occurrence of earthquake and flood during the
service life of bridge is potentially low. However, certain regions are more prone to both
seismic and flood hazard, such as region close to snowy mountain and fault line. Therefore, it
is critical to account for the possibility of earthquakes and floods that occur concurrently in
these regions. Several studies are performed in past to investigate the joint vulnerability of
bridge structures which are exposed to earthquakes and flood-induced scour (Alipour and
Shafei 2012, Alipour et al. 2013, Banerjee and Ganesh Prasad 2013, Yilmaz et al. 2016). The
seismic vulnerability model of river-crossing bridges which are exposed to flood scenario is
expressed probabilistically using fragility function that are commonly represented as

𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑜𝑓 𝑏𝑟𝑖𝑑𝑔𝑒 𝑢𝑛𝑑𝑒𝑟 𝑐𝑒𝑟𝑡𝑎𝑖𝑛 𝑑𝑎𝑚𝑎𝑔𝑒 𝑠𝑡𝑎𝑡𝑒 = 𝑃(𝐷𝑆𝑘 | 𝑅𝑃𝐹𝑙𝑜𝑜𝑑 ) (2.1)

where, RPFlood represents the recurrence interval of flood hazard which is a commonly
adopted parameter in design of bridges. Eq (2.1) represents the probability that a bridge exceeds
15
or meets a specific damage state for a given intensity measure (IM) of earthquake and
recurrence interval flood event. Generally, the fragility curves are expressed as a two-parameter
log-normal distribution function (Banerjee and Shinozuka 2007), for which the median value
is denoted as ck and measured in terms of peak ground acceleration (PGA) and the log-normal
dispersion is represented as 𝜁𝑘 . These parameters indicate, respectively, the level of PGA for
which there is 50% chance of exceeding a particular damage state k (such as minor, moderate,
major and collapse) and dispersion of fragility curves. Mathematically, the analytical form of
the fragility function is expressed as

  PGAj 
 ln  
F ( PGAj , ck ,  k ) =    k 
c
(2.2)
 k 
 
 

Ganesh Prasad and Banerjee (2013), and Banerjee and Prasad (2013) generated fragility
curves for different span length of multi-span concrete box girder bridges in California using
non-linear time history analysis method. The fragility curves of considered bridges are
generated based on Eq. (2.2) considering scour depth around bridge foundation as intensity
measure for flood hazard using the methodology proposed in Shinozuka et al. (2000).
Consequently, Yilmaz et al. (2016) adopted flood return period as an IM instead of expressing
scour depth as IM in seismic fragility curves of river-crossing bridges. Wang et al. (2014a)
developed vulnerability model for various types of bridge classless which were commonly
found in United States. These include seismically designed single-frame concrete box girder
bridges with integral piers, older non-seismically designed multi-span simply supported
(MSSS), and multi span continuous (MSC) concrete girder bridges. It is observed that as scour
depth increases, the probability of column failure in box-girder bridges increases. Interestingly,
in the case of MSSS and MSC bridges, scour can have positive impact on seismic response of
piers. This is due to the presence of scour which results in a longer vibration period of bridges,
similar to the scenario of base isolation. Wang et al. (2014b) conducted a similar study and
generated fragility surface for complete damage state for the same type of bridges in California.
Both studies (Wang et al. (2014a, 2014b) utilized probabilistic seismic demand model (PSDM)
that uses regression analysis to establish a correlation between engineering demand parameters
(EDP) with chosen IM of seismic hazard. Multi-hazard fragility function involving earthquake
and flood is also developed by utilizing reliability methods and Bayesian networks (Gehl and
D’Ayala 2016).

16
Past research has also examined the impact of earthquake, flood-induced scour, and
corrosion on the safety and integrity of bridges. In a study by Li et al. (2020), the time-
dependent seismic performance of two-span bridges was analyzed under the influence of scour
and chloride-induced corrosion. The findings indicate that as the bridge ages, the flood-induced
scour and corrosion can weaken the bridge isolation provided by laminated elastomeric
bearing, leading to pier damage during earthquakes. The literature has also investigated the
effect of both corrosion and carbonation on joint probability of bridge failure under the
interaction of earthquake, scour and corrosion (Ranjkesh et al. 2019), and proposed a retrofit
solution to enhance the bridge resistance against these potential hazards.

Previous studies have examined the scour depth around bridge foundations either
deterministically (Ganesh Prasad and Banerjee 2013) or used probabilistic approaches to assess
the performance of bridges under seismic hazard for various flood return periods (Yilmaz et al.
2016). However, these investigations did not take into account climate change as a factor for
bridges that are susceptible to scour. Khelifa et al. (2013) attempted to evaluate the failure risk
of bridge under climate change induced scour by extrapolating the intensity of 100-year flood.
Similarly, Dong and Frangopol (2016) employed the same methodology to evaluate the seismic
performance of scour-vulnerable bridges in the context of climate change. To explore the
impact of climate change on failure probability of bridges, Kallias and Imam (2016) modified
the dataset for river discharge by adjusting its mean and standard deviation, and thus
established a correlation between two variables. Uncertainties associated with the future
projections of precipitation cannot be accounted by conventional approach such as altering
mean and standard deviation of 50, 100 and 200-year flood (Anderson et al. 2015). Only the
results of global climate model (GCM) and downscaling techniques are suitable methods to
derive regional level data that quantifies the flood hazard considering various climate related
scenarios (Khandel and Soliman 2019). Yang and Frangopol (2019) proposed a framework that
utilize GCM and downscaled simulation data to predict future climate. Hydrologic modelling
is used to convert climate simulation data into flow discharge data to quantify scour. To account
for uncertainties in future climatic conditions, the approach incorporates multiple climate
models and climate futures.

2.2.2 Vulnerability model of bridges under Level-II interaction: Earthquake and corrosion

The interaction of earthquake and corrosion is categorized in Level -II multi-hazard


interaction, which has the potential to cause compounded effect on bridge structures through

17
direct or indirect interaction. In the past, several studies have assessed the influence of chloride-
induced corrosion of embedded rebars in pier on seismic vulnerability of highway bridges
(Choe et al. 2008, Ghosh and Padgett 2010, Shivang et al. 2018). These studies demonstrated
the importance of accounting corrosion deterioration in the assessment of seismic fragility of
bridge structures. The significance of considering both uniform (Deng et al. 2018, Vishwanath
and Banerjee 2019) and pitting (Ghosh and Sood 2016) form of corrosion in vulnerability
assessment of bridge structures were highlighted in these studies. The seismic vulnerability
model of bridge classes which are commonly found in Central States of United States (CSUS)
were derived for pristine bridges in Neilson (2005). Subsequent research efforts by Ghosh and
Padgett (2010) and Ghosh (2013) have focused on incorporating the effect of corrosion into
the assessment of seismic vulnerability over the lifespan of CSUS bridges. Ghosh and Padgett
(2010) developed fragility curves for MSC steel girder and MSSS concrete girder bridges
(Ghosh 2013) that account for the impact of chloride-laden deicing salts on the steel
reinforcement of columns and bearing over time. In addition to chloride exposure on piers,
Simon et al. (2010) investigated the seismic vulnerability of RC bridges in California by
considering the effect of corrosion deterioration on piers and as well as on pile. Shivang et al.
(2018) also considered exposure scenarios such as marine splash and atmospheric zone to
assess the life-time seismic vulnerability and seismic loss of RC highway bridges. Rao et al
(2017) developed time-dependent fragility function for degraded bridges columns that were
designed during three distinct design periods, demonstrating that the vulnerability of bridges
reduced with the improvement in design standards.

Although significant efforts have been made in the literature to evaluate the impact of
corrosion-induced deterioration on seismic vulnerability of highway bridges, majority of these
studies only accounted for the scenarios of deterioration in the substructure of the bridge. The
upper portion of bridge comprising deck, girder and other load bearing members is also
susceptible to gradual deterioration. This scenario potentially leads to a decrease in structural
integrity of a bridge, compromising its ability to withstand loads, including seismic forces. The
measurement of surface chloride concentration of 15 bridge decks over a period of 15 years
indicates that the measured chloride concentration exceeds the allowable limit in reinforced
cement concrete (Kassir and Ghosn 2002). The interaction of earthquake, corrosion, and traffic
hazard has not been extensively explored in the literature. However, Zhu and Frangopol (2013)
developed a computation approach to quantify the time-dependent risk caused by traffic and
earthquake loads. The study indicated the significance of accounting traffic load in determining

18
the failure probability of each girder in a bridge deck, which in turn affects the overall system
failure probability of the bridge. Decò and Frangopol (2013) further explored the risk-
assessment of aging bridges by developing a framework that integrates bridge deterioration,
seismic hazard, and traffic hazard model to estimate the total risk of bridges. These two studies
proposed a framework that can be useful for decision-makers to prioritize bridge maintenance
and rehabilitation activities to reduce vulnerability of bridge over its life-cycle. Few studies
have proposed a framework to evaluate the life-time seismic performance of corroded cable-
stayed bridges (Lu et al. 2018, Pang et al. 2022).

2.3 Definition of limit-states in the literature to develop fragility curves

In the view of generating fragility curves, a limit state is defined in the literature which
refers to a threshold at which a structural or non-structural component is expected to reach a
specified level of damage or performance due to earthquake or seismic loading. For example,
a limit state of bridge pier may be referred as the state at which the pier exceeds a certain drift
ratio or suffers a certain degree of cracking. Several limit state thresholds for bridge piers are
suggested in the literature in terms of concrete and rebar strain, drift, displacement, and
curvature ductility. Similarly, for abutments and bearings under seismic load, the threshold
limit is based on its deformation in the longitudinal and transverse directions. After the 1994
Northridge earthquake in California, Basoz and Kiremidjian (1998) developed a classification
system that categorized damage threshold for bridges into various states such as minor,
moderate, major and collapse. The qualitative description of such damage state are mentioned
in HAZUS-MH (2009). In the early 1990’s, Priestly et al. (1996) defined concrete crack width
and steel strain as limiting criteria to define limit state. It was suggested that concrete cracks
should remain small, not more than 1mm, so that it can categorized as minor damage. To
indicate higher level of damage, it was suggested that the rebar tensile strain and concrete
compressive strain should not exceed 0.015 and 0.004. Material strain cannot be easily read by
field inspectors after earthquakes. As a result, researchers explored alternative indicators that
can be easily quantified through instrument aided observations.

In earthquake engineering, drift ratio and curvature ductility of piers are commonly
adopted as limit state threshold to categorize damage state of piers under seismic loading. By
definition drift ratio is defined as the ratio of lateral displacement of pier to its height. It is a
quantitative measure of deformation that a pier undergoes during a seismic event. As its ratio
increases, the damage suffered by the pier also increases. Berry and Eberhard (2003)

19
established regression equation that correlate drift ratio with longitudinal buckling of rebar (∆𝐿𝑏𝑏)
∆𝑠𝑝𝑎𝑙𝑙
and concrete cover spalling ( 𝐿
). These equations were obtained from experimental
observations of 300 rectangular columns and 170 circular columns. Sheikh and Yau (2002)
indicated drift ratio for different magnitude of axial loads. Table 2.1 provides the threshold
limit state of cylindrical pier by other researchers which are quantified in terms of drift ratio.
Similarly, threshold limits of elastomeric bearing in terms of deformation in the longitudinal
and transverse directions, and abutment with backwall height hw are listed in Table 2.2 and
Table 2.3.

Curvature ductility is also adopted to categorize damage states of piers when subjected
to seismic loading. By definition curvature ductility is the ratio of maximum curvature of a pier
during seismic loading to the curvature corresponding to the yielding of reinforcement. It
measures the ductility or ability of the pier to sustain deformation before failure. Seismic design
has undergone significant improvement since 1971, resulting in an enhanced improvement in
the ductility of columns. Prior to 1971, columns designed with older methods were typically
brittle and lacked ductility. From 1971 to 1990, seismic provision improved, resulting in
columns designed for enhanced ductility that was considered as strength degrading under
extreme lateral load. Post-1990, columns have experienced significant improvement in
ductility and are considered as ductile columns. Curvature ductility of columns along with
damage states for columns designed from 1971 to post 1990 are documented in Table 2.4. The
experimental results related to performance of bridge columns are well documented in Mackie
and Stojadinovic (2005) and Veletzos et al. (2006).

2.4 Performance-based design criteria for highway bridges

Performance-based design (PBD) of highway bridges employs limit-state threshold as


a quantitative criterion. This approach was first developed in New Zealand during the 1970s
and after it evolved in the United States. It has been widely adopted as a design provision
globally, including in the Canadian highway bridge design code (CSA 2014) and the New
Zealand bridge design manual (NZ 2016). In the last two devastating earthquakes namely the
1994 Northridge and the 1995 Kobe earthquake, several bridges incurred losses as a result of
post-event disruptions in the aftermath of seismic event, although many bridges survived the
seismic event without collapse. Hence, it became crucial to ensure that the seismic design of
bridges should not only ensure seismic safety of bridges but also confirms continuous
serviceability of bridges in the highway network.

20
Table 2.1 Limit state thresholds of cylindrical piers in terms of drift

Reference Threshold limit Damage state Other details

Drift (%) = 1.0 Minor

Dutta and Drift (%) = 2.5 Moderate Drift limits for seismically
Mander (1998) Drift (%) = 5.0 Major designed bridges

Drift (%) = 7.5 Collapse

Drift (%) = 0.7 Slight

Kim and Feng Drift (%) = 1.5 Moderate Drift limits for non-seismically
(2003) Drift (%) = 2.5 Extensive designed bridges

Drift (%) = 5.0 Collapse

- Minor P = axial load,

∆𝑠𝑝𝑎𝑙𝑙 𝑃 𝐿 fck = concrete strength, L = height


(%) = 1.6 (1 − ) (1 + 0.10 ) Moderate
𝐿 𝑓𝑐𝑘 𝐴𝑔 𝐷 of column, ke = coefficient for
Berry and
∆𝑏𝑏 𝑘𝑒 𝜌𝑒𝑓𝑓 𝑑 𝑃 𝐿 column shape, Ag = gross area of
Eberhard (2003) (%) = 3.25 (1 + ) (1 − ) (1 + 0.10 ) Major
𝐿 𝐷 𝑓𝑐𝑘 𝐴𝑔 𝐷 section, D = diameter of column,
𝜌𝑒𝑓𝑓 = longitudinal reinforcement
Collapse
ratio

21
Reference Threshold limit Damage state Other details

- Minor

∆𝑠𝑝𝑎𝑙𝑙 𝑃 𝐿
(%) = 1.6 (1 − ) (1 + 0.10 ) Moderate
𝐿 𝑓𝑐𝑘 𝐴𝑔 𝐷 𝜌𝑙 = longitudinal rebar ratio,

Mackie and ∆𝑏𝑏 150𝜌𝑒𝑓𝑓 𝑑 𝑃 𝐿 𝜌𝑡 = traverse rebar ratio.


(%) = 3.25 (1 + ) (1 − ) (1 + 0.10 ) Major
𝐿 𝐷 𝑓𝑐𝑘 𝐴𝑔 𝐷
Stoiadinovic
(2007) ∆𝑐𝑜𝑙
(%) = 16 + 0.71. 𝐿 − 1.5𝐷 Similar to Berry and Eberhard
𝐿
(2003)
− (3.8 × 10−5 𝑓𝑐𝑘 × (−2.8 × 10−5 ). 𝑓𝑦 ) Collapse
𝑃
+ (1.7𝜌𝑙 + 1.9𝜌𝑡 ) − (12 )
𝑓𝑐𝑘 𝐴𝑔

Drift (%) = 0.6 % Minor Drift ratio for cylindrical pier in


Drift (%) = 2.5 % Moderate transverse direction of bent
Kwon and
Elnashai (2010) Drift (%) = 4.2 % Major

Collapse

22
Table 2.2 Limit state thresholds of elastomeric bearings
Reference Threshold limit Damage state Other details
Shear deformation (%) = 150 Minor Yielding of steel plates
Shear deformation (%) = 106-150% Moderate Yielding of anchor bolts
Mori et al. (1997)
Shear deformation (%) = 200 Major Bearing uplift
Shear deformation (%) > 300 Collapse Bearing failure
Longitudinal and transverse deformation (mm)= 30 Minor Minor cracks
Longitudinal and transverse deformation (mm)= 100 Moderate Yielding of anchor bolts
Nielson (2005)
Longitudinal and transverse deformation (mm)= 150 Extensive Fracture of anchor bolt
Longitudinal and transverse deformation (mm)= 255 Collapse Unseating
Longitudinal and transverse deformation (mm)= 40-85 Minor Tension failure of anchor

Konstantinidis et Longitudinal and transverse deformation (mm)= 86-309 Moderate Initiation of slipping
al. (2008) Longitudinal deformation (mm)= 310-532 Major -
Longitudinal deformation (mm) > 532 Collapse Unseating
Shear deformation (%) = 53-85% Minor Tension failure of anchor
Shear deformation (%) = 100-200% Moderate Initiation of slipping
- Major -
LaFave et al.
Shear deformation (%) > 400 % Collapse Unseating
(2013)

23
Reference Threshold limit Damage state Other details
Shear deformation (%) = 20 Minor Potential yielding of anchor
Shear deformation (%) = 100 Moderate Yielding of shim plates
Stefanidou and
Kappos (2017) Debonding between rubber
Shear deformation (%) = 200 Major
and shim plates
Shear deformation (%) = 300 Collapse Unseating

Table 2.3 Limit state thresholds of abutments

Reference Threshold limit Damage state Other details


- Minor -
- Moderate -
Romstad et al. Passive wedge failure in
(1995) Longitudinal deformation (mm) = 0.03.hw Major
backfill soil
Ultimate deformation of
Longitudinal deformation (mm)= 0.1.hw Collapse
abutment-backfill
Longitudinal deformation (mm) = 0.003.hw Minor Minor cracks in backwall
Longitudinal deformation (mm) = 0.006.hw Moderate Yielding of backfill soil
Choi (2002)
Longitudinal deformation (mm) = 0.06.hw Major Ultimate deformation
- Collapse -

24
Reference Threshold limit Damage state Other details
Longitudinal deformation (mm)= 50-100 Minor Minor cracks in backwall
Reduction of abutment
Shamsabadi et al. Longitudinal deformation (mm) > 100 Moderate
stiffness
(2007)
- Major -
- Collapse -
- Minor -
- Moderate -
Wilson and Passive wedge failure in
Elgamal (2008) Longitudinal deformation (mm)= 0.047.hw Major
backfill soil
Ultimate deformation of
Longitudinal deformation (mm)= 0.10.hw Collapse
abutment-backfill
Longitudinal deformation (mm)= 1.10.dgap Minor Minor cracks
Longitudinal deformation (mm)= 0.01.hw Moderate Yielding of abutment soil
Stefanidou and
Kappos (2017) Deformation of abutment
Longitudinal deformation (mm)= 0.035.hw Major
soil
Longitudinal deformation (mm)= 0.10.hw Collapse Ultimate deformation

25
Table 2.4 Limit state thresholds of cylindrical piers in terms of curvature ductility

Curvature ductility
Year of design Damage state
(𝝁𝝋 )

0.80 Minor

0.90 Moderate
Pre 1971 (Brittle column)
1.00 Major

1.20 Collapse

1.00 Minor

2.00 Moderate
1971 – 1990 (Strength degrading column)
3.50 Major

5.00 Collapse

1.00 Minor

4.00 Moderate
Post 1990 (Ductile column)
8.00 Major

12.00 Collapse

26
2.4.1 Basics of performance-based seismic design framework

A flow-chart for performance-based seismic design of bridges is illustrated in Fig. 2.3. The
first step in PBD involves conducting a comprehensive hazard analysis to evaluate one or more
intensity measures (IMs) associated with ground motions. At the given location (O) and for the
bridge facility (D), IM is determined using traditional probabilistic seismic hazard analyses and is
generally expressed as the mean annual probability of seismic hazard (denoted as p[IM]). Each
standard design provision such as AASHTO and other bridge design codes proposed distinct IM
for seismic hazard. For instance, AASHTO (2017) recommends a seismic hazard of IM
corresponding to 1000-year return period. In contrast, Eurocode (2005) recommends 475-year and
95-year return period seismic events, while the Canadian highway bridge design code (CSA 2014)
suggests 2475-year and 475-year return periods seismic events for conducting safety assessment.

Following the steps in the figure, structural simulation must be performed for the bridge
facility under the selected ground motion to acquire the engineering demand parameters (EDP).
EDPs are used to represent the response of the structure in various ways such as deformations,
forces, and accelerations. For bridges, the most commonly used EDPs include curvature ductility
of piers, longitudinal and transverse deformations of abutments and bearings. These performance
simulations are performed using finite element models (FEMs) of representative/existing bridges.
The following step in the framework is to perform damage analysis which establishes a correlation
between EDPs and DM, thereby representing the actual physical damage incurred by the facility.
The final step in the framework is to quantify decision variables, DV, that are quantified in terms
of monetary consequences. These variables are expressed in many alternative ways such as single
scenario (Fig. 2.4a), continuum of scenarios (Fig. 2.4b), loss with particular probability of
exceedance (Fig. 2.4c), and loss with different level of exceedance (Fig. 2.4d).

2.4.2 Existing research on multi-hazard risk assessment of bridges in the context of PBD

In literature, the risk is considered as a performance indicator to determine whether the


structural performance of a bridge under extreme loading is termed as acceptable or unacceptable
(Ayyub and Popescu 2003, Ellingwood 2005). Therefore, PBD approaches consistently
incorporate the consideration of risk. Given the importance of assessing multi-hazard in PBD for

27
bridges, existing literatures on risk assessment involving multi-hazard will be taken into account
for the review. Wen (2001) presented a framework for designing structures taking into account the

Figure 2.3 Flow-chart of PBD of bridges

Figure 2.4 Different forms of expressing seismic risk of bridges

minimization of life-cycle cost under multi-hazard such as earthquakes and winds. A similar study
by Han and Frangopol (2023) adopted life-cycle risk as performance indicator to identify optimal
maintenance solution for bridge networks that are susceptible to corrosion and seismic hazard.
Decò and Frangopol (2011) developed a comprehensive framework for evaluating risk associated
28
with highway bridges that are exposed to a range of stressors, including traffic loads, scour caused
by flood hazard, and earthquake. Kameshwar and Padgett (2014) indicated a considerable variation
in the relative risk associated with hurricane and earthquake hazard with different range of pier
height. The study demonstrates a non-linear reduction in hurricane risk and linear increase in
seismic risk as the height of column increases. A similar study by Markogiannaki (2019) found
that bridges exposed to earthquake and wind hazard had a lower risk due to seismic hazard. Yilmaz
(2015) documented an extensive knowledge base on risk assessment of highway bridges that are
exposed to multi-hazard involving earthquake and flood. The study also conducted an uncertainty
analysis to quantify the variability of risk. Shivang et al. (2018) evaluated a life-cycle seismic risk
of highway bridges under various exposure conditions such as deicing salt, atmospheric and
marine splash scenarios during its service period. In contrast, Zhong et al. (2023) developed a
framework that quantifies life-cycle seismic risk for entire bridge lifespan, including both
construction and service period.

Although there have been multiple published studies on the topic of risk assessment of
bridges under multi-hazard events, there are still literature gaps that require attention, particularly
in the context of PBD of bridges. Firstly, there is a need to develop a comprehensive framework
that examines the impact of pier design parameters on the seismic performance of bridges with
flood-induced scour, considering different design level earthquakes. Second, while most existing
studies have focused on specific region like CSUS and Europe to access hazard risk for bridges,
there is a necessity to explore the multi-hazard risk of bridges for other developing regions, such
as the high-seismic and flood-prone region of Asian countries. Finally, most studies have not
incorporated climate change-related stressors into risk assessment of bridges that are exposed to
multi-hazard events involving earthquakes and flood. Hence, there exists a need to perform study
to address such literature gap.

2.5 Life-cycle deterioration of embedded rebar in girder under the influence of corrosion-
fatigue

RC bridges are designed and built to be functional for long service time, but the mechanism
of corrosion and fatigue can reduce their performance over time. When the embedded rebar in
concrete comes in contact with aggressive substances like chlorine, carbon, and other chemicals,
it will undergo corrosion. This process dissolves the material of rebar, leading to the formation of

29
rust, which in turn reduces its effective cross-sectional area. Fatigue occurs when highway bridges
is subjected to repeated loads from truck traffic, which results in development of short cracks
arising from the location of pits and progress to long crack over time. The synergetic action of
corrosion and fatigue significantly accelerates the deterioration process over time and reduces its
structural performance. The life-cycle deterioration of embedded rebar in RC girder due to
corrosion-fatigue is divided into four stages for discussion and each stage is illustrated in Fig. 2.5.

• Stage I: The degradation of embedded rebar in RC girders begins after construction due
to infiltration of chloride ion through various sources such as deicing salt, atmospheric
and marine splash scenarios. This process is known as diffusion which leads to the
dissolution of protective film around the rebar and initiates corrosion. The time when
corrosion starts is referred to as the time of corrosion initiation (tini).
• Stage II: The primary mode of deterioration of rebar in this stage is corrosion. At this
stage, corrosion results in formation of pit on the surface of rebar which provide location
for the fatigue cracks to initiate due to repeated loading.
• Stage III: This stage is primarily dominated by fatigue. Over time, due to repeated loading
or stress, micro-cracks arise from the location of pits, which grow and propagate faster
than the corrosion rate, and lead to reduction in the efficient cross-sectional area of rebar.
The number of cycles of repeated loading that cause the micro-cracks to develop and
propagate until the rebar fail by fracture is referred as fatigue life of rebar.

The fatigue life of rebar is determined experimentally by subjecting the rebar to cyclic
loading and monitoring its number of cycles which causes failure. The outcome is typically plotted
on an S-N curve, which illustrates the relationship between stress amplitude (difference between
maximum and minimum stress level) and the number of cycles to failure. The S-N curve is
generally used to estimate the fatigue life of rebar under specific loading conditions.

2.5.1 Existing fatigue life prediction model of corroded steel rebar

Past studies have explored to formulate a mathematical model that attempts to predict the
fatigue life of rebar under cyclic loading and that are exposed to aggressive environment. The
literature has identified two primary methods to predict the fatigue life of rebar, ‘total life
approach’ and the ‘damage-tolerance approach’.

30
Figure 2.5 Dissolution of embedded rebar in girder due to corrosion-fatigue

The concept of total life approach investigates the fatigue life of metal samples under both
inert and aqueous condition. The result of this experimental observation is generally represented
by Stress-Life (S-N) or Strain-Life (ε-N) curves. Depending upon the number of cyclic loading or
fatigue life, the fatigue failure is classified as high cycle fatigue (HCF) and low cycle fatigue
(LCF). If the number of cyclic loads exceed 104 but remain within 106 on metal sample, it is
referred to as HCF. In contrast, if the number of repetitive loads is within 104, it is classified as
LCF. Generally, bridge girders are subjected to HCF with low-stress level (Higgins et al. 2007).

Generally, when fatigue test is performed in metal samples, the observation is typically
represented in terms of number of load cycles (or fatigue life Nf) that the specimen can withstand
prior to fracture, against the applied stress range Δσ. This mathematical relationship is expressed
as

𝑁𝑓 (Δσ)𝑚 = 𝐴′ (2.3)

where m and 𝐴′ are exponent and material constant that can be derived based on fatigue test, and
the value of m generally varies between 3 and 5. This curve is generally termed as single-slope S-
N curve that exhibits a linear relationship between stress range and the logarithm of number of
cycles to failure. This means that the fatigue life of the material decreases exponentially as the
stress range increases. The stress-life relationship of metal sample under HCF was initially

31
modelled empirically by Basquin (1910). The stress range Δσ is related to fatigue life Nf of metal
with fatigue strength coefficient 𝐴′ and exponent m is expressed as
𝛥𝜎 𝑚
= 𝐴′ (2𝑁𝑓 ) (2.4)
2

Eq. (2.4) is applicable to metal samples which are subjected to HCF, when metal samples
is subjected to LCF it results in plastic deformation. As a result, the material can only withstand
few numbers of cycles. Since the deformation is no longer elastic, the fatigue behaviour of the
metal is defined by the strain range. The stress-life of the metal samples is mathematically
modelled by Manson (1953) and Coffin (1954) and it is expressed as
𝛥𝜀𝑝 𝑝
= 𝐵 ′ (2𝑁𝑓 ) (2.5)
2

where, 𝛥𝜀𝑝 represents the plastic strain range and 𝐵 ′ and b refers fatigue ductility and exponent
coefficient. The above expressions (i.e., Eq. (2.4) and Eq. (2.5)) are derived using sinusoidal
loading. In fatigue analysis, the mean stress is the average of stress over one loading cycle. In the
case of sinusoidal loading, the mean stress is zero. However, in practical scenario, mean stress of
the fatigued sample varies around non-zero value, and hence the applicability of Eq. [2.4] and Eq.
[2.5] have its limitation.

The concept of damage tolerance approach is also known as fatigue crack propagation
approach which provides more realistic solution for fatigue problems in RC structures, thus
overcoming the limitations associated with conventional total life approach. It utilizes linear
elastic fracture mechanics (LEFM) to evaluate the growth of cracks under applied external stress.
LEFM is a technique adopted in literature for studying the fatigue behaviour of materials
containing pre-existing cracks. This method involves performing stress analysis in the area
surrounding the crack to monitor the propagation of crack until it exceeds its critical size. The
stress intensity approach is commonly adopted in the literature to investigate the propagation of
cracks under cyclic loading, based on the fundamental assumption that each stress component
induced by the crack growth is directly proportional to stress intensity factor (SIF).

2.5.2 Existing the impact of corrosion-fatigue in RC structures: Experimental approaches

There exists an ample number of experimental observations on corrosion fatigue in RC


structures. In a study conducted by Ahn and Reddy (2001), an experimental investigation was

32
performed to assess the durability of RC beam under fatigue loading and chloride ingress. The
study performed experimental loading on 16 RC beams and considered the impact of static and
cyclic loading with varying water/cement ratio (such as 0.3, 0.4 and 0.6). Marine tidal exposure
was simulated to induce accelerated reinforcement corrosion. The observation of beam specimen
under 78000 cycles of loading are shown in Fig. 2.6, and the result indicate that beam specimen
subjected to cyclic loading during the exposure period exhibited reduced ultimate strength as
compared to those subjected to static loading. The fatigue load which applied in the study by Ahn
and Reddy (2001) are fixed pulsating loads which involves the application of constant cyclic load
that is repeated over a period of time, while experimental study with moving loads involves the
application of load that moves along a predetermined path, which simulates the real-life condition
of bridge structures that are subjected to fatigue. Experimental test was performed in past on small-
scale bridge model under the moving load to investigate the fatigue failure mode of RC slab (Okada
et al. 1978; Perdikaris and Beim 1988; Maekawa et al. 2006). Numerous experimental
investigations have been conducted on RC beams to explore the impact of corrosion and fatigue
on their structural performance (Zhang et al. 2017; Li et al. 2019).

Generally, these test setup focus on either the independent effect of corrosion or fatigue
loading, whereas the combined effect of corrosion and fatigue are not accounted for in the test
setup. However, past studies have investigated the combined interaction of corrosion and fatigue
on the performance of RC rectangular beam (Lu et al. 2018) and T-girder specimens (Zhang et al.
2023). Micro-structural analysis methods such as scanning electron microscopy (SEM) and X-ray
diffraction are used to study the microstructural changes in rebar specimens caused by corrosion-
fatigue damage. Blikharrsky et al. (2021) used electronic microscopy to analyze reinforcement
samples and found that samples exposed to aggressive environment for 15-30 days shows
corrosion pits up to 60-70 µm in size, which led to the development of fatigue cracks under cyclic
load. Lu et al. (2022) determined that both internal defects of the base metal and corrosion-induced
pitting on the surface of rebar are the primary causes of fatigue failure. Yaohab et al. (2022)
observed that cracks responsible for corrosion fatigue failure originated from various locations on
the specimen surface and that mechanical damage played a significant role in the initiation of
fatigue cracks under high-stress conditions.

33
Figure 2.6 Ultimate strength for RC beams under static and cyclic loading

Experimental studies have shown that corrosion reduces the fatigue life of RC beams. Sun
et al. (2015) developed an empirical correction coefficient for flexural stiffness of corroded RC
beams under fatigue load. Ma et al. (2014) found that concrete fatigue degradation has little effect
on fatigue life of RC beam, and suggested a simple fatigue life model that accounts corrosion-
fatigue degradation in steel rebar. Ma et al. (2018) observed that the flexural stiffness of corroded
RC beam degrades with an increase in fatigue load.

2.5.3 Existing the impact of corrosion-fatigue in RC structures: Analytical approaches

Analytical approach is also adopted in the literature to assess the fatigue performance of
bridges. This involves developing mathematical model of bridge and simulating the passage of
vehicles and performing computer simulations to predict the behaviour of bridges under moving
load. In the past, analytical methods such as finite element analysis (FEA), stress-life and strain-
based, and fracture mechanics-based approaches are adopted in the literature to assess the fatigue
performance of bridges. FEA is a computer simulation technique which is adopted to analyze the
structural behavior of bridges under different loading conditions. FEA can be utilized to model the
component of bridges and predict its stress and strain that occur under repeated moving loading.
Heshamti (2012) utilized structural hot spot stress and the effective notch stress methods for the
fatigue assessment of steel bridges. Most of the studies on stress-life method followed reliability-

34
based fatigue analysis of bridges. Byers et al. (1997) have published an in-depth review on
approaches for fatigue reliability of steel bridges in highways and railways. When using stress-life
method for fatigue reliability, the two key factors to consider are fatigue load and resistance. To
determine the fatigue load mode, it is important to consider not only its magnitude but also its
frequency of occurrence. One way to obtain this information is through weigh-in-motion (WIM)
measurements. On the other hand, the resistance model should be based on the sample which was
tested with significant number of fatigue loading conducted under variable amplitude load. WIM
measurement can provide accurate data on variables such as number of axles, velocity of vehicle,
axle weight and GVW. This information can then be used as input parameters to develop fatigue
load model. Additionally, various methods have been proposed in the literature to derive the
fatigue strength distribution such as the approach given in Murthy et al. (1995). Zhao et al. (2000)
have also developed a method to determine an approximate distribution of fatigue life when limited
amount of test data is available. Yan et al. (2017) proposed a computational methodology to derive
fatigue stress spectra for short-to-medium span bridges utilizing WIM measurement. The study
utilized stress-life approach to determine the time-varying fatigue reliability of T-girder RC bridge
for various level of traffic growth. Lu et al. (2021) used the similar framework and stress-life
approach to derive fatigue reliability of aging prestressed concrete bridges accounting for the
stochastic natures of traffic load and reinforcement corrosion.

Structures which are subjected to low-cycle fatigue are assessed for fatigue performance
through strain-life method. In this method, strains measured in bridge component are considered
to be plastic rather than elastic. Researchers have conducted studies to evaluate the fatigue
performance of steel bridges using theoretical strain-life method (Sakano and Wahab 2001; Ge et
al. 2013). However, there exist only a few number of studies that use strain-life method to assess
the fatigue life of steel bridges, particularly in RC bridges, where most of the fatigue related
problem are associated with high cycle fatigue.

Fracture mechanics approach is also adopted in literature to assess the fatigue performance
of bridge structure. This approach utilizes various parameter such as SIF, initial crack size, and
fracture toughness to characterize the behavior of crack under different loading conditions, and
forecast whether the crack will propagate and result in failure or not. Several studies have been
carried out using fracture mechanics approach to assess the fatigue reliability of bridges (Rocha
and Brühwiler (2012); Bastidas-arteaga et al. (2014); Accornera et al. (2022)). Guo and Chen
35
(2011) suggested a reliability based LEFM model for a 40-year-old steel-box girder bridge, taking
into account of uncertainties in stress data. Bastidas-arteaga (2009) developed probabilistic
fracture-mechanics based lifetime prediction model for RC T-girder which are subjected to
corrosion-fatigue. Pipinato et al. (2011) adopted a probabilistic approach with LEFM to evaluate
the fatigue reliability of steel bridges under seismic loading.

2.6 Vertical ground motion and corrosion-fatigue bridges: An essential connection

In the previous section, it is indicated that highway bridges are susceptible to damage
caused by corrosion-fatigue, and result in reduction of flexural stiffness of bridge deck. Therefore,
it is imperative to undertake seismic fragility assessment to determine the most suitable retrofit
measure that can enhance the structural performance of bridge during seismic event. The effect of
vertical ground motion (VGM) is commonly characterized by a response spectrum that derived
from horizontal ground motion (HGM) after reducing its spectral acceleration with a factor two-
third (Newmark et al. 1973). However, observation from past seismic events such as 1989 Loma
Prieta, 1994 Northridge, 1999 Chi-Chi, and the 2001 Bhuj earthquake in India reveals that the
spectral vertical acceleration can exceed the suggested acceleration level for a certain period of
time (Elgamal and He 2001). According to various studies on this topic (Amraseys and Simpson
1996; Collier and Elnashai 2001), the ratio of peak horizontal ground acceleration to peak vertical
ground acceleration has been underestimated for a region in near-faults. Furthermore, these studies
indicated the spectral acceleration ratio of HGM to VGM exceed the commonly used value of two-
third at periods shorter than 0.2 sec, particularly at near-source distances less than 30 km.
Additionally, it is also commonly observed from past seismic events that the frequency content of
VGM is quite different from HGM. When considering bridge structures with corrosion-fatigued
damage and that had lost its flexural stiffness during their service life, the impact of VGM is more
complex. The impact of VGM on bridge structures that are in deteriorated state due to corrosion-
fatigue has not been examined in previous studies. Therefore, the relationship between VGM and
the condition induced by corrosion-fatigue on bridge structure during seismic events must be
examined to fully understand the impact on deteriorated bridge structures that have lost their
flexural stiffness due to corrosion-fatigue.

36
2.6.1 Existing studies for investigating the effect of VGM on performance of highway bridges

Saadeghvaziri and Foutch (1991) conducted pioneering investigation to explore the impact
of VGM on highway bridges. The study performed experimental test on RC column and observed
that the VGM produced varying axial force on the column, causing the rebar or column to yield
under a moment that a potentially below the anticipated design capacity. Papazoglou et al. (1996)
conducted an analytical investigation and compared the obtained result with field damages. It was
indicated that some modes of bridge failure can traced back to the impact of VGM. Additionally,
it was also indicated that the fluctuation in axial force would result in both shear and flexure failure.
Veletzos et al. (2006) performed a study on the seismic behavior of precast segmental joints of
bridge superstructures using a suite of 10 near-field earthquake records, and indicated that the
application of VGM has resulted in a 400% increase in the median positive bending moment of
bridge deck. Moreover, Gullerce et al. (2012) has investigated the EDP such as positive and
negative bending moment of girder at the face of bent cap, and positive and negative bending
moment of girder at the middle of span in the perspective of VGM, and found out that the inclusion
of VGM would result in degrading performance of bridges. This finding is further supported by
the analytical investigation conducted by Wibowo and Sridharan (2022). Kim et al. (2011)
examined the effect of VGM on bridge pier. The study explored the variable V/H (the ratio of PGA
in the vertical direction to that in horizontal direction) with different range of magnitude and the
time lag between the arrival of peak horizontal and vertical acceleration. It was observed that the
increase in V/H ratio resulted in a significant increase in axial force within the pier, potentially
reducing the shear capacity by as much as 30%. While the arrival time had a minimal impact of
axial force variation and shear demand, but it had a notable impact on shear-capacity.

The nonlinear-time history analysis (NLTHA) of a representative short-to-medium span


bridge located in Washington reveals that the seismic event with V/H ratio exceeding one can
cause vertical uplift of the superstructure at the abutment and could result in damage incurring due
to pounding (Wilson et al. 2015). It was indicated that the VGM is potentially vulnerable to bridges
in region that are prone to moderate earthquake. The study determined a design PGA of 0.32g for
the case-study region. However, Shrestha (2015) presented analytical evidence that the ground
motion with highest PGA in the vertical direction dos not necessarily result in maximum relative
response. The amplification of response is primarily influenced by the V/H ratio of the seismic
event, rather than PGA of HGM. Prior research has shown that VGM can induce fluctuating axial
37
forces in the column. However, these studies have not observed the impact of VGM on bearing
and the resulting variations of its normal force. This is an important consideration as the change in
normal force can alter the load transfer mechanism of bridge system. Several investigations have
indicated that the normal forces on the bearing during VGM can vary substantially, and in some
cases, even it develops tensile forces in the bearings for the bridges which are close to near-fault
(Warn and Whitaker 2008; Marin-Artieda and Whittaker 2010). A series of experimental
investigation by Xian and Li (2017) on elastomeric bearing subjected to VGM indicates that
coefficient of friction between the elastomeric and surface of the girder varies in a nonlinear
inverse relationship with normal stresses.

The seismic performance of individual components of highway bridges (such as the girder,
column, bearing and abutment) under liquefaction in the presence of VGM was investigated by
Wang et al. (2013). Al-Attraqchi et al. (2022) studied to investigate the utilization of concrete-
filled steel tube (CFST) columns as a sustainable alternative for withstanding both horizontal and
vertical ground motions, with a focus on increasing resilience. Khakzand et al. (2022) performed
an analytical investigation on simply supported three-span bridge subjected to near-fault ground
motion with horizontal, vertical, point and cord components. Their finding revealed that the
response of bridge to a combination of HGM and VGM was identical to its response to horizontal-
vertical and point-cord ground motion components. Despite having extensive literatures on
investigating the effect of VGM on highway bridges, there is still a lack or limited of research on
the seismic fragility assessment of bridges subjected to combined HGM and VGM. However,
Wang et al. (2013) developed a fragility model for various components of bridges, including
girder, bearing and pier, in the context of liquefaction-prone bridges under the influence of VGM.
The study emphasized the importance of considering VGM in both demand and capacity model
for developing fragility model. Thapa et al. (2022) developed a fragility model for a typical bridge
in Nepal at the system level. The study performed seismic vulnerability analysis that compared the
vulnerability model under the effect of horizontal-only and simultaneous consideration of HGM
and VGM excitation on bridges.

Several studies have attempted to investigate the rationality of the code and standardized
provision that govern bridge design in the context of VGM. For example, Button et al. (2002)
performed parametric investigation to assess the effect of VGM on critical components of standard
highway bridges. The study indicated that the most significant impact is felt by bridges whose
38
vertical period matches with the maximum spectral acceleration in the vertical response spectrum.
The authors suggested to discontinue the practice that utilizes the vertical spectra with amplitude
equals to two-third of horizontal spectra, particularly for the bridges which are situated close to
the fault line. Another, Kunnath et al. (2008) conducted simulation to investigate the seismic
response of two-span highway overcrossings. The results showed that the vertical component of
ground motion caused a significant increase in axial force demand in column and moment demand
in girder both at midspan and face of the bent cap. The observation revealed that negative moment
in the midspan section exceeds its capacity. This indicates that seismic design criteria, such as
Caltrans SDC (2019) which only accounts for vertical effect through a static load equivalent to
25% of dead load, need to be revised.

2.7 Seismic resilience of highway bridges

The concept of resilience is widely adopted as performance indicator in infrastructure


engineering. It is well known fact that the earthquake can cause significant damage to bridges,
disruption to functionality to transportation system, that results in huge economic loss. Given the
large number of bridge structures in the infrastructure portfolio of developed and developing
countries, repairing bridges that suffer severe earthquake damage can be difficult due to the
requirement of high cost. This highlights the inadequacy of designing structures based solely on
life safety principles. Instead, seismic design should also consider economic and social impact of
structural damage. Therefore, the literature has introduced resilience-based design principles to
address this issue. By considering the potential hazard and consequences of multi-hazard event in
the design process, resilience-based design aims to improve the ability of bridge to resist, adapt
and quickly overcome from the disaster event. This approach helps to ensure that the bridge can
continue serving its intended purpose, prevent long-term economic consequences, and reduce the
need for costly repair and replacement.

The resilience characterizes the ability of bridge system to resist and absorb the impact of
disasters event along with the vulnerability, robustness, risk and adaptability of system. The term
‘resilience’ was first coined in the field of biophysics and ecology (Holling 1973). The definition
of resilience is quoted as ‘Resilience is a magnitude of disturbance which a system can tolerate
and still persist’. After the introduction of resilience, the concept was utilized and extended to

39
many branches of science and engineering, and the concept was even found its application in a
recent pandemic of COVID-19 (Prime et al. 2020).

2.7.1 Estimation of seismic resilience through quantitative approaches

The generic numerical expression to quantify the seismic resilience of civil infrastructure
system using quantitative approach was suggested in several studies (Bruneau et al. 2003; Henry
and Ramirez-Marquez 2012; Cimellaro 2016). Bruneau et al. (2003) presented a framework that
is widely recognized and established to evaluate resilience of civil infrastructure system, with
specific emphasis on seismic hazard. Seismic resilience of highway bridge refers to the structural
capacity of bridge to withstand the effect of earthquake and swiftly recover from its disruption to
restore its intended function. A conceptual representation of the measurement of resilience through
quantitative approach is depicted in Fig. 2.7. The figure illustrates a scenario in which the
infrastructure undergoes a disaster event at time t0, resulting in a decline in its functionality that
potentially leads to the closure of traffic lanes. The magnitude of the reduction in functionality
depends on the vulnerability of infrastructure (higher robust system leads to lower vulnerability).
The recovery process begins at t0 and is completed by time tr, with the speed of recovery
represented by the slope of the functionality curve between these two points. The recovery rate is
influenced by both the level of damage and the availability of resources. Once the recovery process
is complete and the infrastructure functionality has been restored to pre-disaster level, its resilience
is evaluated at control time tc. The magnitude of resilience for the disaster event that occurred at
time to is calculated by determined the area under the functionality curve between to and tc.

Robustness, rapidity, redundancy and resourcefulness are four main properties which
characterize the resilience. Robustness refers to the ability of bridge system to withstand a
disruptive event without failing or breaking down. Rapidity refers to the progress which a system
can recovery from a disaster event. A system that can quickly return to pre-event performance after
subjecting to disruption is more resilience than one takes a long time to recover. Resourcefulness
denotes the ability of bridge system or network to quickly mobilize resources in response to
seismic event to restore the intended functionality of bridge. Redundancy refers to the presence of
additional system or components that can be used if the primary system fails.

40
Figure 2.7 Conceptual illustration of seismic resilience

The seismic resilience R of a bridge system from the time of seismic incident t0 up to a
control time t0, Bocchini and Frangopol (2012) have suggested a formulation that involves
determining the area under the functionality curve Q(t) (Fig. 2.7) within the time frame bounded
by t0 and t0+T and it is given as
𝑡 +𝑡𝑐 𝑄(𝑡)
𝑅 = ∫𝑡 0 𝑑𝑡 (2.6)
0 𝑇

𝑄(𝑡) = 1 − [𝐿(𝐼, 𝑇𝑟𝑒𝑐 ) × {𝐻(𝑡 − 𝑡𝑂𝐸 ) − 𝐻(𝑡 − (𝑡𝑂𝐸 + 𝑇𝑟𝑒𝑐 )) × 𝑓𝑟𝑒𝑐(𝑡, 𝑡𝑂𝐸 , 𝑇𝑟𝑒𝑐 }] (2.7)

Here L(I,Trec) is a loss function comprising of direct and indirect losses, and frec(t,tOE,Trec)
is recovery function which will be discussed in further subsequent section and H(•) is the Heaviside
step function. Using the quantitative approach several studies assessed the seismic resilience of
bridges under single and multi-hazard scenarios (Bocchini et al. 2014; Chandrasekaran and
Banerjee 2016; Deco et al. 2013; Dong and Frangopol 2015; Karamlou and Bocchini 2015;
Venkittaraman and Banerjee 2014; Zheng et al. 2018; Vishwanath and Banerjee 2019; Devendiran
et al. 2021; Devendiran and Banerjee 2023). These studies utilized the conventional bridge fragility
model to represent the vulnerability of bridges under hazard scenarios. Direct and indirect losses
are derived from the bridge fragility, and each study adopted a suitable recovery model that
considered factors such as level of damage of infrastructure, the availability of resource, and the
capacity of the stakeholder to respond and recover.

41
2.7.2 Estimation of seismic resilience through qualitative approaches

The literature also adopted qualitative assessment procedure to estimate bridge resilience.
These procedure aid in quick assessment of resilience for the presence of large number of bridges
which exists in infrastructure portfolio. A number of studies, such as Ikpong and Bagchi (2015),
Domaneschi and Martinelli (2016), Andric and Lu (2017) and Minaie and Moon (2017) have
utilized qualitative approach in their investigation aimed at simple decision-making procedure for
computing the resilience of bridges in US, Canada, and China. Ikpong and Bagchi (2015) and
Domaneschi and Martinelli (2016) developed methods to compute resilience index (RI) of
highway bridges in response to extreme climate change events. In the procedure, weight factors
are utilized to consider bridge replacement cost (as a percentage of initial construction cost),
consequences of extreme event and post-disaster cost. To quantify RI, weight factors for various
bridge components are multiplied by corresponding bridge ratings against capacity measures. The
resulting products are then summed across as identified resilience indicators to provide RI for a
given bridge. Andric and Lu (2017) discusses the use of fuzzy logic in quantifying the seismic
resilience of bridges. The study suggests that utilizing fuzzy logic is a valuable method for
addressing uncertainties that may arise due to insufficient data. Minaie and Moon (2017) suggested
a qualitative method for evaluating the robustness of bridges to multiple hazards that accounts four
key factors, namely hazard, importance of bridge, vulnerability and uncertainty in recovery time.

2.8 Post-event losses and recovery model for quantifying resilience

Bridge failure results in direct and indirect post-event losses. The direct loss includes the
cost of repairing and replacing the bridge in the aftermath of disaster event. Meanwhile, indirect
loss includes business interruption, time-loss, environmental damage, and other factors which are
attributed due to the damage of bridge. Quantifying resilience requires consideration of each loss
and recovery strategies, and it is location specific.

2.8.1 Existing direct loss models

When a bridge fails during a seismic event, there are various cost of expenditure arise to
stakeholder for the repair and restoration process. This cost can be categorized into three groups
(i) reconstruction and rehabilitation (ii) cost required to remove debris (iii) cost required to provide
temporary by-pass. These group of cost is together termed as direct loss, which can be obtained by

42
multiplying the bridge area (width and length) by the sum of restoration cost associated with each
possible outcome (Shinozuka et al. 2005). Most of the published literature on the topic adopt the
direct loss estimation model, which originally suggested in HAZUS (HAZUS-MH 2004). This
model determines the direct loss (denoted as CD) using fragility curves resulting from bridge
damage and post-event repair by multiplying the probability of bridge damage denoted as
PE(DS=k) at different damage state k with their respective damage ratio rk with the total cost of
bridge replacement (C). The analytical expression for calculating the CD is written as follows

𝐶𝐷 = ∑4𝑘=1 𝑃𝐸(𝐷𝑆 = 𝑘). 𝑟𝑘 . 𝐶 (2.8)

where, rk is taken as 0.03, 0.08, 0.25, and 0.67, respectively for k = 1–4 (HAZUS-MH 2004) and
information on bridge vulnerability PE (DS = k) are obtainable from bridge fragility curves. The
visual representation of direct loss estimation is illustrated in Fig. 2.8. This approach is begun by
selecting the appropriate IM for the bridge (e.g., PGA or PGV) (step 1), and then determine the
probability of damage exceeding a certain damage state, such as minor, moderate, major and
collapse (step 2). Multiply this probability by a designated damage, rk (step 3), and then calculate
the direct loss for the particular IM by multiplying the total cost of bridge replacement (step 4).

Decò et al. (2013) modified the expression given in Eq. (2.8) and included costs associated
with bridge rehabilitation/reconstruction, removal of debris, and construction of a temporary
bypass, and it is expressed as

𝐶𝐷 = 𝑊𝐿 ∑𝑁{𝑃𝐸 (𝐷𝑆 = 𝑘)[𝛼𝐶𝑟𝑒𝑏 𝑑𝑘 (1 + 𝑉𝑐 ) + 𝛽𝐶𝑟𝑒𝑚 + 𝛾𝐶𝑟𝑒𝑏 𝑏𝑘 (1 + 𝑉𝑐 )]} (2.9)


where, W and L are width and length of bridge deck in m, N represent the number of
recovery options available in a decision tree to repair a bridge after damage, Creb and Crem are,
respectively, unit area rebuilding (i.e., reconstruction) cost and debris removal cost ($/m2), Vc is
the velocity coefficient that related to recovery pace, bk is the proportionality factor between cost
for construction of temporary bypass and bridge replacement value, and α, β, γ are the coefficient
related to the outcomes of recovery action decision tree. Here, reconstruction cost is calculated by
assuming it to be proportional to the bridge replacement value as is considered in HAZUS
(HAZUS-MH 2004). In a similar manner, costs for construction of a temporary bypass and debris
removal are assumed to be proportional to bridge replacement value and deck area, respectively.

43
Figure 2.8 Direct loss estimation procedure using fragility curve

Mackie et al. (2016) suggested a different approach by computing direct loss through the
evaluation of the expense required to fix the damaged bridge components, which were then
calculated based on the anticipated carbon footprint using the following numerical expression.

𝐿(𝑖𝑚) = 𝑅𝐶𝑅(𝑖𝑚)[1 + 𝛼𝐶𝐹𝑅 𝐿𝐶𝐹𝑅 (𝑖𝑚)] (2.10)

in which, L(im) represent the loss function associated with bridge damage, while RCR is
the repair cost ratio, which can be estimated by considering all repair items from all performance
group (i.e., bridge components related to repair work). To assess sustainability, penalty function
αCFR and LCFR is used, which measures the ratio of carbon footprint of repair to that of new
construction. Although the later two models (Eqs. (2.9) and (2.10)) are more comprehensive than
the HAZUS model (Eq. (2.8)); however, the HAZUS model is more popular due to its simplicity
and minimal input requirement.

2.8.2 Existing indirect loss models

For estimating indirect losses, the literature identified three different approaches (i) the use
of traffic data for calculation (ii) rough and approximate calculations, and (iii) no estimation. Two
separate model based on traffic data have been proposed in the literature. One of these models,
introduced by Stein et al. (1999) accounts traffic data and costs associated with detours (Decò et
al. 2013; Dong and Frangopol 2015; Dong and Frangopol 2016; Zheng et al. 2018; Viswanath and

44
Banejee 2019). The model calculates indirect loss by adding the operating cost of vehicles on
detour (Cop) and cost due to vehicle time loss (CTL) owing to bridge after the event. The operating
cost of vehicles (Cop) due to detour is expressed as

(2.11)

where Cop,car and Cop,truck are the average costs of operation of car and truck per kilometre length
($/km), respectively, Dl is the detour length (km), ADT is the average daily traffic, and TRD is the
average daily truck traffic ratio (%). Cost due to time loss (CTL) of vehicle users and goods
travelling through detours can be given as

(2.12)

where cost terms CAW, CATC and Cgoods are average wage per hour ($/h), average total compensation
per hour ($/h), and monetary value of time taken to transport goods in cargo ($/h), respectively.
Ocar and Otruck are average vehicle occupancies for car and truck, respectively, l is the length of
link (km), S is the average speed on detour (km/h), SD and S0 are average speeds on the damaged
and intact bridge (km/h), respectively, and ADE is average daily traffic remaining on the bridge
after the event. Values of these parameters are case specific, and should be taken based on the
bridge under consideration. For an example, state Departments of Transportation (DOTs) and the
Federal Highway Administration (FHWA) in the U.S. annually report travel demand on bridges;
hence, bridge-specific ADT values can be obtained directly from these sources.

The other traffic data-based indirect loss model considers that a new bridge replaces the
post-event damaged bridge and no potential redistribution of traffic occurs on alternative routes
(Bocchini et al. 2014). Daily indirect loss (Cind) for bridge damage due to a seismic event (E) is
expressed as

(2.13)

45
where ADToverpass cars and ADToverpass trucks are average daily cars and trucks on overpasses,
respectively, and ADThw cars and ADThw trucks are the same on crossed highways. loverpass and lhw are
additional lengths of detour respectively for the overpass and crossed highway (in miles), sa is the
average speed on the detour in mph, ctime cars and ctime trucks are cost of time for cars and trucks,
respectively (USD/h), and cdistance cars and cdistance trucks are cost per mile covered by cars and trucks,
respectively (USD/m). To obtain total expected indirect cost, Cind is multiplied with the
probabilityof occurrence (PE) of the seismic event over bridge life cycle and then the product is
integrated over bridgerecovery period following the recovery path.

In the absence of such traffic information, approximateestimation of indirect loss is done


by taking it to be an order of magnitude higher than direct loss for seismically damaged bridges
(Chandrasekaran and Banerjee 2016; Venkittaraman and Banerjee 2014). Such a simple model can
be used only when ground motions are expected to produce considerable damage to bridges that
would give rise to a significant indirect loss (i.e. the ratio of indirect to direct loss is high). Hence,
difficulty persists in developing an appropriate model for indirect loss. Such complex formulations
of indirect loss estimation are avoided in Karamlou & Bocchini (2015) and Mackie et al. (2016),
and bridge resilience is estimated there by taking direct loss only.

2.8.3 Recovery model for assessing resilience

Indirect loss persists in the damaged network until the system is recovered to its pre-event
functionality level. Based on the expert opinion survey data (ATC 1985), HAZUS (FEMA 2003)
developed recovery functions for highway bridges at minor, moderate, major damage and
complete collapse states. These recovery functions take normal cumulative distribution functions
for all damage states, along with their respective distribution parameters, including mean (µ) and
standard deviation (σ). For minor damage state, both µ and σ are set to 0.6 days. The moderate
damage state has µ and σ values of 2.5 and 2.7 days, respectively. The major damage state is
characterized by µ and σ values of 75 and 42 days, while the collapse damage state has µ and σ
values of 230 and 110 days, respectively. A number of past studies on resilience assessment of
bridges used this model to simulate post-event recovery (Dong and Frangopol 2016; Zheng et al.
2018).

While using the HAZUS model as a base, different mathematical forms such as linear,
trigonometric and exponential are also used to simulate the path of post-event recovery of a real-

46
life bridge and results are compared to judge the relative suitability of these functions. It was
observed that linear recovery function can provide a reasonably good and realistic estimate in the
absence of post-event recovery information, and hence is used in several recent studies on
resilience assessment of bridges (Chandrasekaran and Banerjee 2016; Mackie et al. 2016; Minaie
and Moon 2017). Bocchini et al. (2014) used the survey data from ATC (1985) and formed
stepwise functional recovery in terms of traffic flow capacity of bridges throughout their recovery
periods. Stepwise function recovery is also proposed by (Sharma et al. 2018), in which the steps
involve three essential phases of recovery – planning, execution and closure. It is worth mentioning
here that the HAZUS model is associated with uncertainties that may eventually lead to high
variability in computed resilience.

Vishwanath and Banerjee (2019) proposed sigmoidal function (also called as a logistic
function) as a recovery model to represent the recovery path of damaged bridges (illustrated in
Fig. 2.9). The function is controlled by two parameters (a and b) that influence its shape and make
it capable of producing different recovery paths depending on the nature of bridge damage,
probable time to mobilize recovery actions and time to complete recovery process.
Mathematically, the bridge recovery function frec can be expressed in terms a sigmoidal recovery
function as

1
f rec (t ) = (2.14)
1 + e − a ( t −b )

where t represents a time instant, and a and b are two parameters mentioned earlier. As shown in
Fig. 2.9, the rapidity of a recovery curve is influenced by a and the inflation point b controls the
movement of the curve along x-axis, and the study evaluated these shape parameters a and b in
terms of recovery time Trec.

The expected value of Trec at a damage state is taken to be equal to the time at which 84%
functional recovery is achieved at the same damage state from FEMA (2003) specified restoration
curves for highway bridges. By doing so, a full correspondence is established between mean +
three standard deviation value of Trec at a damage state with HAZUS-specified time for 100%
functional recovery at the same damage state. Thus, expected values of Trec equal to 1.2, 5.2, 117
and 340 days are obtained at minor, moderate, major damage and collapse states of bridges,

47
respectively, and used for deriving recovery curves (Fig. 2.9). Corresponding standard deviation
values are 0.35, 1.9, 24.2, and 63 days, respectively.

(a) (b)

(c) (d)
Figure 2.9 Recovery models at different bridge damage states; (a) minor damage; (b) moderate
damage; (c) major damage; and (d) collapse state

A generalised multi-parameter sinusoidal-based continuous function is proposed by Deco


et al. (2013) to model bridge recovery after seismic events. The six model parameters are residual
functionality, idle time, recovery duration, target functionality and two shape parameters. The
proposed model is used by the authors in their subsequent studies on the same topic (Dong and
Frangopol 2015). The formulation assumes recovery duration to follow a triangular distribution.
The proposed formulation can capture recovery paths according to post-event damage condition
of bridges (i.e. in the forms of stepwise, linear, positive and negative exponential functions and

48
sinusoidal curves). Evidently, this model is more complex when compared to the one prescribed
in HAZUS; however, calibration of model parameters may pose a great challenge if appropriate
post-event recovery related information from real-life scenarios is not available. For bridges in a
network (real and virtual), total costincurred due to bridge damage during a natural disaster is
obtained according to adopted post-event restoration models for bridges in the same network.
Taking real and virtual highway transportation networks, Bocchini and Frangopol (2012a; 2012b)
proposed a restoration model based on idle time and speed of restoration. It is assumed that the
time taken for restoration has linear dependency on the damage level of bridges. Other studies
related to seismic resilience ofbridge networks used the recovery model proposed in HAZUS
(HAZUS-MH 2004) with or without case-wise customisations (Alipour and Shafei 2016;
Karamlou and Bocchini 2016). Readers can also refer to Gidaris et al. (2017) for a discussion on
available restoration models of highway bridges for the purpose of resilience assessment under
earthquake and tsunami hazards.

2.9 Literature gap

After reviewing the existing literatures in the focus area of this dissertation, the following
research gaps were identified and summarized below. These gaps in the research define the scope
of the current study.

• Understanding how natural hazard interact with each other and impact the life-cycle
performance of river-crossing bridges would improve the preparedness of the stakeholder
prior to extreme events and that could ultimately benefit the society. A quantitative
framework which integrates the effect of climate change on multi-hazard performance of
bridges involving earthquake and flood is yet to addressed in the literature for assessing
the risk and resilience of river-crossing bridges. In addition, climate change adaptative
measures for addressing issues related to the multi-hazard threat should also be identified
and its economic benefit should be determined through cost-benefit analysis;
• For river-crossing bridges, the literature needs to address the impact of design parameter
related to pier, such as aspect ratio, longitudinal reinforcement ratio, and differential
elevation between multiple pier bents, on the multi-hazard performance of these bridges
under earthquake and flood scenarios. It is necessary to introduce a performance-based

49
design framework to facilitate the probabilistic risk assessment and design of river-
crossing bridges under this multi-hazard scenario.
• For bridges which are located in urban areas with dense traffic flow and corrosive
environment, corrosion and traffic-induced fatigue are possible external stressor that
affects the life-cycle seismic performance of bridges. A quantitative framework to
effectively evaluate the material degradation of RC girders under the combined interaction
of corrosion and fatigue needs to be developed to aid in assessing the seismic performance
and resilience of highway bridges over their lifespan.
• For such bridges, girders lose flexure stiffness due to corrosion-fatigue degradation.
Hence, dynamic characteristics of these bridges alter over their lifespan. Therefore, it is
necessary to examine the impact of VGM on such deteriorated bridges and determine how
their seismic vulnerability changes throughout service life.

2.10 Closure

This chapter explores various multi-hazard conditions for bridges, discusses multi-hazard
impact on seismic fragility and risk of bridges, and identifies potential sources of threats over
lifecycle of bridges located across rivers and in urban areas. It discusses performance indicators,
such as risk and resilience, for assessing bridge performance under multi-hazard threats, and the
methodology to evaluate such indicators. Existing research gaps related to life-cycle based
resilience assessment of highway bridges are highlighted. Describing the complexity of multi-
hazard conditions, the chapter emphasizes the need for developing a holistic approach for bridge
design and maintenance throughout its service life.

50
This page is intentionally left blank

51
CHAPTER 3

MULTI HAZARD PERFORMANCE OF RIVER-CROSSING BRIDGES IN


CHANGING CLIMATE: RISK, RESILIENCE AND ADAPTATION

Climate change attributes to more frequent and severe floods, which will affect the
infrastructure, particularly river-crossing bridges. Flood events can increase the velocity and
discharge of river, causing erosion of riverbed, and results in higher scour around bridge piers.
This higher scour can make the bridge substructure more flexible and unstable, posing a threat to
seismic safety and functionality of the bridge. Current bridge design does not fully consider the
potential impact of climate change, which could result in negative consequences for the
community in the long run. To address this issue, this chapter conducts a study using available
data for the California region to derive future flood projections using a general circulation model
under changing climate condition. The chapter also proposes an integrated approach to assess the
impact of climate change on bridge performance in the context of risk and resilience under multi-
hazard threat involving earthquake and flood. The study focuses on a real-life existing bridge in
San Joaquin River in California and suggests an adaptation strategy for future disaster multi-hazard
event. The chapter also highlights the importance of incorporating climate change consideration
into the design and maintenance of river-crossing bridges to improve the resilience and decrease
the seismic risk in long run.

3.1 Integrated approach for the adaptation of bridges to changing climate

The adaptation of river-crossing bridge to the changing climate in this study was addressed
by following the similar procedure suggested in Lund University (2018). The procedure consists
of four primary steps. First, the problem is identified. Second, the impact of the problem is assessed
in terms of structural vulnerability. Third, the vulnerability is evaluated and expressed in monetary
terms. Finally, a suitable adaptation measure is identified that can reduce the vulnerability of the
structure under changing climate condition improve the resilience and reduce the risk of the bridge
in the aftermath of disaster event. These steps are illustrated in Fig. 3.1.

52
Figure 3.1 General methodology to identify the adaptation solution for climate change

It is believed that the operational status of the bridge is at risk due to four significant
changes which would potentially bring by climate change (also illustrated in Fig. 3.2). These
changes include (1) rise in frequency and severity of flooding event which could result in scouring
around bridge foundation (2) rise in temperature and frequent occurrence of heat waves that results
in additional stress on bridge components, impaired functioning of expansion joints (3) increase
in concentration of anthropogenic CO2 and relative humidity that might enhance the change of
concrete carbonation and hasten the corrosion (4) rise in sea level which would potentially
submerge coastal bridges. Generally, to address the impact of climate change on bridges, a general
approach is adopted in literatures. First, climate data and historical records of precipitation,
temperature changes are gathered for the case-study region. Subsequently, an assessment is
performed out to determine the vulnerability of bridge to various factors such as scouring,
thermally induced stresses etc. Following that, an analysis of the cost-effectiveness of the
adaptation measure is conducted. Therefore, this problem should be addressed in multi-
disciplinary approach which requires expertise in various scientific field including climatology,
hydrology and hydraulics, structural engineer for risk and life cycle cost assessment.

53
Figure 3.2 Projected changes in environment and its impact on bridges due to climate change

3.2 Short description of case-study bridge

To demonstrate the adaptation solution for river-crossing bridges under changing climate,
the selected case-study bridge is located on Interstate 5 in San Joaquin County, California.
Interstate 5 is the longest interstate in California stretching from the Mexican to the Canadian
border near Washington. Given the importance of truck traffic to the agricultural industry in the
region, this interstate serves as an important transportation route. As a result, the case-study
bridge was required to facilitate traffic to cross across the San Joaquin River. Fig. 3.3 illustrates
the pictorial aerial view of the case-study bridge.

The bridge has six-spans of prestressed concrete box-girder bridge with a total length of
235.3m long and was constructed in 1972. The girders are monolithically connected to wall-type
piers, and each pier is founded on a group of prestressed concrete piles with a reinforced concrete
pile cap. The bridge is firmly laid on foundation which consist of silty sands. The bridge has
integral abutments on both sides and an in-span hinge between Bent 3 and Bent 4. At the hinge,
the bridge girder is provided with separation of 25.4 mm (or 1-in.) gap such that one side of the
girder sits on elastomeric bearings placed on the other side of the girder. The elevation view of the
bridge is shown in Fig. 3.4. The San Joaquin River is prone to flooding, particularly in winter and

54
early spring, when heavy rainfall and snowmelt can cause the river to overflow its banks. The river
has a history of flooding, with several major incidents over the last century. During the most recent
bridge inspection in 2015, the National Bridge Inventory (NBI) item code 113 (scour criticality
code or, SCC) of the bridge was identified to be 5, which explains that bridge foundations are
stable for evaluated scour condition and scour is within footing limit (NBI 2015). This rating is
just at threshold; SCC below 5 signifies that either the bridge is scour critical or necessary actions
should be taken to protect exposed foundations from further erotion (FHWA 1995). Hence, it is
very appropriate to study the probable projection of scour at bridge piers during future intensive
flood events in that region owing to the climate change.

Figure 3.3 Aerial view of case-study bridge (picture curtesy: Google Map)

Figure 3.4 Elevation view of case-study bridge

55
3.3 Finite element model of the case-study river crossing bridge

The numerical simulation of the case-study bridge is performed in open-source finite


element program called OpenSees. Three-dimension (3D) finite element model (FEM) is generally
used to simulate the behavior bridge structures under extreme loading events. The 3D FEM in
OpenSees is used to represent the material, geometry and boundary condition of the case-study
bridge. The development of 3D FEM in OpenSees entails a series of stage-by-stage steps for
abutments, girders, piers, and foundations. All of these components are represented as nodes and
are linked together with the appropriate element. Each element represents mass, stiffness, and
damping of the associated components. The schematic representation of FEM of the bridge is
shown in Fig. 3.5.

3.3.1 Prestressed girders

In the FEM, prestressed concrete girders is modelled as a one-dimensional beam element


that represents the mass, stiffness and damping of the girder. Since, the bridge girders are expected
to remain elastic during a seismic event, the structural behavior such as shear, moment and axial
load are idealized linearly. In this study, elasticbeamcolumn elements are assigned to model bridge
girders in OpenSees. These elements are assigned to the center of gravity (CG) of the section, and
the mass associated with that element is lumped on the nodes to which they are connected. Since
the girder is prestressed concrete, the flexural rigidity of the element is assigned based on the gross
sectional property of the girder (Yilmaz et al. 2016).

3.3.2 Piers (wall-type)

Displacement-based fiber (DBF) elements are generally used in the finite element analysis
(FEA) of flexural components including piers. The versality and compatibility of such DBF
elements with wide range of linear and nonlinear dynamic analysis makes these as a preferred
choice for FEA. These fiber-based elements simulate the structural behavior of a pier by
discretizing into patches of fiber with appropriate constituent materials (such as concrete and
steel). This fiber adds strength and stiffness to the section. The fiber section of wall-type pier is
shown in Fig. 3.5. To quantify displacements and forces, the element employs principles of virtual
work. In OpenSees, Concrete07 and Steel02 are assigned to fiber for defining the material model

56
of concrete and steel. The stress-strain relation of confined and unconfined concrete of wall-type
pier is assigned based on material model suggested by Mander et al. (1988).

Figure 3.5 Schematic representation of 3D finite element model

3.3.3 Pile foundations

The FEM of pile is represented using elasticbeamcolumn, and the properties associated
with the element are assigned based on the sectional properties of piles. The interaction between
the pile foundation and the surrounding soil is idealized by placing a series of springs
equipped with nonlinear materials that characterize the seismic resistance of soil at different depths
of the pile foundation. The lateral resistance of pile-soil interaction is idealized using PySimple1
and the shaft resistance of pile-soil interaction is modelled using TzSimple1 material in OpenSees.
Both PySimple1 and TzSimple1 are used in conjunction with pile element to model the soil-
structure interaction behavior of pile foundation in the lateral and vertical directions of the bridge.
The strength and deflection relation (p-y or t-z) for PySimple1 and TzSimple1 are assigned based

57
on the suggestions of American Petroleum Institute (API 2003) and Mosher (1984). Expected
scour depth caused by a flood event (discussed in the following section) is calculated and modelled
in FEA by removing soil springs from the existing ground elevation to the level of scour depth.
Following that, strength-deflection properties of remaining springs are re-calculated and updated
in FEM.

3.3.4 Abutments

The bridge has integral abutments on both sides. These abutments are casted monolithically
with the bridge girder. This type of design offers several benefits which includes reduced
maintenance and improved seismic performance. In the FEM, the backwall-backfill interaction of
abutment is idealized using elastic-perfectly plastic backbone curve as suggested in SDC (2019).
The force-deflection relation of abutments is represented using ElasticPP material in OpenSees,
and it is shown in Fig 3.6a. The passive-force of the soil imposed on the abutment (Pabut) is
calculated as given in Caltrans (2019) and expressed below.

𝑤 ℎ
𝑃𝑎𝑏𝑢𝑡 = 𝐴𝑒 × 239 𝑘𝑃𝑎 × 1.7 (𝑚, 𝑘𝑁) (3.1)

Here, Ae and hw are effective height and area of abutment wall. The units of these
parameters should be entered in meters (m). The stiffness of the back-bone curve illustrated in Fig
3.6b is calculated as given in SDC (2019).


𝑎𝑏𝑢𝑡
𝐾𝑎𝑏𝑢𝑡,𝑙 = 𝐾𝑖 × 𝑤 × 1.7 (3.2)
𝑚

where, w is the projected width of diaphragm wall, and the initial stiffness of the material is given
as 28.7 kN/mm/m for embankment fill material that meets the requirements of Caltrans Standard
Specifications. The transverse response of integral abutment of case-study bridge is modelled as
per the recommendation suggested in Aviram et al. (2008). ElasticPP material is utilized to
idealize the behavior of wing-wall and pile. The strength and stiffness of the back-bone curve of
the wing wall with height hww and with effective width weff in the transverse direction is calculated
by modifying the Eq. (3.2) by accounting effectiveness factor (CL = 2/3) and participation
coefficient (Cw = 4/3)
ℎ𝑤𝑤
𝑃𝑎𝑏𝑢𝑡,𝑡 = 𝐶𝑤 × ℎ𝑤𝑤 × 𝑤𝑒𝑓𝑓 × 239 𝑘𝑃𝑎 × 1.5 × (𝑚, 𝑘𝑁) (3.3)
1.7

58

𝐾𝑎𝑏𝑢𝑡,𝑡 = 𝐶𝑤 × 𝐾𝑖 × 𝑤𝑒𝑓𝑓 × 1.7𝑤𝑤𝑚 (3.4)

, ,

,
,

Figure 3.6 Back-bone curve of abutment in (a) longitudinal (b) transverse direction

3.3.5 Elastomeric bearing at in-span hinge

The structural response of elastomeric bearing in the translational direction is idealized


using material which have linear elastic perfectly plastic back-bone curve, and this material is
assigned to zerolength springs in two orthogonal horizontal directions. The shear stiffness (Kshear)
and shear strength (Fy) of back-bone curve is computed as

𝐴×𝐺
𝐾𝑠ℎ𝑒𝑎𝑟 = (3.5)
𝑇𝑟

𝐹𝑦 = 𝜇 × 𝑁𝑏𝑒𝑎𝑟𝑖𝑛𝑔 (3.6)

where, G represents the shear modulus of the elastomer, A represents the plan area of elastomer, 𝜇
is the friction coefficient between elatomer and surface of girder, and 𝑁𝑏𝑒𝑎𝑟𝑖𝑛𝑔 is the axial force
due to gravity loading.

3.4 Regional flood hazard and its impact under changing climate (observed & projected)

The flood hazard curve at a given region represents the likelihood of flooding at a specific
location. These are typically developed through statistical analysis of historical annual peak flow
of a given river. The historical annual peak flow data which are observed from 1930 to 2011 at the
nearest gauging station of case-study bridge in San Joaquin river (USGS Station #11303500) is

59
shown in Fig 3.7. These data are publicly available on the National Water Information System
(2013). Snowmelt runoff from the Nevada mountains is a major source of water for the San Joaquin
River, especially during the spring and early summer months. The river receives a significant
amount of run-off following the melt of snowfall that occurred during the winter (Welty & Zeng
2021). Future projections under global greenhouse gas emissions reveal a possible warming of 1.5-
4.5°C in California by the end of the twenty first century (Cayan et al. 2008). While the frequency
of precipitation events in California may decrease in future, the largest precipitation intensities are
expected to increase (Pierce et al. 2013). Floods in California are generally projected to increase
under climate change (Das et al. 2013). For the San Joaquin River in particular, though total
precipitation amounts aren’t projected to change under global warming, temperatures are expected
to rise with a decline in winter snow-pack, resulting in earlier runoff peaks (USBR 2016). Such
shifts in seasonal cycles can result in increased winter or spring floods associated with earlier
snowmelt (Barnett et al. 2008). The projected discharge of river for the period between 2012-2099
is also shown within Fig. 3.7. These anticipated changes may affect the case study bridge over the
San Joaquin River by increasing its exposure to hydrological events that cause foundation scour
beyond design metrics (i.e., 100-year flood). The background information to quantify the projected
flow between 2012-2099 is documented in the Appendix B of this dissertation.

3.4.1 Flood hazard curve (observed and projected)

Flood hazard curves at the study region are generated through statistical flood frequency
analysis with the annual peak stream flow data over the observed and projected periods due to
changing climate. It is found that the magnitude of projected annual peak flow varies widely over
years,and in some cases, low inflows (<20 m3) were synthesized. Such low peaks are generally
termed as potentially influential points or outliers that have a tendency to change the trend of
remaining data (Klemeš 1986). Hence, elimination of such point sources is necessary for
performing flood frequency analysis to develop flood hazard curves. For addressing the presence
of multiple outliers in the collected flood sample, the updated provision for flood frequency
analysis (Bulletin 17C (USGS 2019)) recommends the multiple Grubbs-Beck test. This guideline
is followed in this study; generated flood hazard curves with annual peak flow data and outliers
are shown in Fig. 3.8 for the observed period (1930–2011) and two non overlapping projection

60
periods, 2012–2050 and 2051–2099. Henceforth, these are referred to as flood Scenarios I, II, and
III, respectively.

Figure 3.7 Annual peak discharge of the case-study river (observed and projected)

Figure 3.8 Mean flood hazard curve for three flood scenarios; Flood Scenario I: Observed, Flood
Scenario II: Projected, and Flood Scenario III: Projected.

61
3.4.2 Flood-induced scour at bridge foundation

In the present study, the 100-year return flood is considered as design flood for the case-
study bridge as it represents a flood event with 1% chance of occurring in any year. This design
flood is generally chosen to ensure that the bridge structure will be able to reasonably withstand
the maximum expected flood event with a reasonable factor of safety (AASHTO 2017). For the
investigated flood event, corresponding discharge from the associated flood event is obtained from
Fig 3.8, and resulting scour-depth around bridge foundation is estimated using the formulation
suggested by Arneson et al. (2012) as shown in the equation next.

𝑎 0.65
𝑦𝑠 = 2.0𝐾1 𝐾2 𝐾3 (𝑦 ) 𝐹𝑟10.43 (3.7)
1

Here, y1 is the flow depth directly upstream from the bridge pier; a is the width of the pier;
K1, K2, K3 are correction factors for pier nose shape, angle of attack of flow, and condition of bed,
respectively. Fr1 is the Froude number which can be obtained by V/√𝑔𝑦1 ; where v and g
represents the velocity of the flow and acceleration of gravity. For the case-study bridge, calculated
scour depth at all bent location is listed in Table 3.1.

Table 3.1 Calculated scour depth at bridge foundation

Type of data Flood scenario

Flood Scenario I (observed) II (projection 1) III (projection 2)

Discharge Q,
3044.46 m3/sec 6224.86 m3/sec 8705.75 m3/sec
(100-yr flood)

Pier Scour at B2 2.53 m 3.02 m 3.27 m

Pier Scour at B3 2.73 m 3.17 m 3.40 m

Pier Scour at B4 2.75 m 3.19 m 3.42 m

Note: where B denotes Bent number of the case-study bridge

3.5 Regional seismic hazard

The case-study bridge in central California is seismically active due to its proximity to the
intersection of the Pacific Plate and the North American Plate, which creates a high risk due to

62
earthquakes. Regional seismicity is accounted for the investigation in two ways in this study. The
first way is to use seismic hazard curves, which exhibit the likelihood of a seismic event occurring
in a specific location over a given time period. The second way is to use regional historic ground
motions for the time-history analysis of the bridge and to develop its seismic vulnerability
model. The development of seismic hazard curve using historic data of available ground motions
of seismic events that occurred near or around the case-study bridge is beyond the scope of this
study. As a result, the seismic hazard curve developed by USGS (2013) for the study region is
taken into account in the analysis here and it is shown in Fig. 3.9a. This hazard curve is used to
assess the financial risk of the case-study bridge for the investigated multi-hazard scenario.

A large dataset of earthquake ground motions with varied intensity of hazard level are
required for the seismic vulnerability of the case-study bridge. The chosen ground motions should
account for the seismic characteristics of the case-study region and should be compatible with the
local soil condition. In this study, recorded ground motions with 170-km radius from the case-
study region is obtained from database of the Pacific Earthquake Engineering Research Center
(PEER 2023). Additional details on selected ground motions are discussed in Yilmaz (2015). To
account for the possibility of future strong earthquakes, some ground motions with lower peak
ground acceleration (PGA) were scaled by a factor of 2. As a result, the final dataset included a
total of 160 earthquake ground motions, with 64, 63, 24, and 9 recordings having PGA values
ranging between 0.1-0.2g, 0.3-0.4g, > 0.4g, respectively. The variability of PGA of the selected
ground motion is shown in Fig. 3.9b.

Figure 3.9 (a) Seismic hazard curve of case-study region (b) PGA of selected ground motion
from PEER database
63
3.6 Change in ductility demand of pier under seismic demand

Nonlinear time history analysis (NLTHA) of the bridge is performed under the chosen
ground motions, and response of critical bridge components are recorded. These response
quantities include post yield flexural response (i.e., curvature ductility 𝜇𝜑 ) of bridge piers,
longitudinal deformations in the active and passive directions at abutments (Δlong,a and Δlong,p,
respectively), and horizontal deformations (Δb,long) at elastomeric bearings (i.e., sliding at concrete-
elastomer interface) in the two orthogonal horizontal direction of the bridge. Curvature ductility
demand is the quantitative measure of the ability of column to deform elastically and plastically
under seismic load. It is obtained as a proportion between curvature that the column experiences
during a seismic event to the yield curvature of the section. Evaluating the deformation of the
abutment and bearing is also equally important as it affects the safety and functionality of the
bridge. Excessive deformation of such components can cause severe damage, resulting in long-
term maintenance and elevated repair costs. As a result, it is critical to evaluate the deformations
of critical components such as piers, abutments, and bearings during seismic events in order to
proportionate the cross-section detail of the components to withstand the expected seismic forces.
These response quantities at critical bridge components define seismic demand at component and
system level of the bridge, and henceforth referred to as engineering demand parameters (EDPs).
Values of EDPs as obtained from NTHA are further categorized in four seismic damage states of
the bridge namely minor, moderate, major damage and collapse states. Physical definition of these
damage states is mentioned and listed in Table 3.2 and quantitative measures for each EDP (i.e.,
thresholds of EDPs at different damage states are summarized in Chapter 2 (refer to Table 2.2 for
elastomeric bearing, Table 2.3 for abutment, and Table 2.4 for pier).

Fig. 3.10 shows the maximum curvature ductility of one of the bridge piers recorded for
all 160 ground motions and for three different baseline periods of 100-year flood. The figure
clearly demonstrates the increasing trend of bridge seismic response as depth of scour increases
due to intense flood events owing to climate change in future years. At some range of peak ground
acceleration (PGA) values (e.g., 0.28g–0.39g), the change in bridge response is significant because
the bridge attains higher damage states as flood scenario changes. This effect is further
demonstrated through fragility curves of the bridge that express vulnerability of the bridge under
the multi-hazard condition.

64
Table 3.2 Damage state definition as given in HAZUS

Damage state Description

Minor Damages that are easy to repair with less expenses.

Significant damage as opposed to minor damage, but still repairable. Before


Moderate
repairs could begin, the bridge structure required temporary bracing.

Damages that are severe and require extensive labour and expenses to
Major complete the repair. Bridges are likely to be shut-down or partial-close down
until the repairs are made.

Significant damage to bridge that cannot be repaired. Bridges in a collapsed


Collapse
state require total reconstruction to restore its intended functionality.

Figure 3.10 Curvature ductility of a bridge pier under varied intensity seismic events for three
flood scenarios

65
3.7 Seismic vulnerability of bridge under changing climate

3.7.1 Component-level fragility curve

The seismic vulnerability of the bridge is represented by fragility curves. These curves are
graphical representations of the likelihood of structural damage or failure during an earthquake
event as a function of PGA or any other intensity measure. Since, PGA is directly correlated and
proportional to the amount of seismic force that can be applied to a structure, it is commonly used
to represent intensity of seismic event in fragility curves. In this study, the seismic fragility curve
is defined analytically using a two-parameter log-normal distribution and it is expressed as
discussed in Chapter 2 (refer Section 2.2.1 and Eq. (2.2)). In the current study, the parameters of
long-normal distribution are calculated by using the method of maximum likelihood, in which the
likelihood function L is expressed as
𝑟
𝑗 1−𝑟𝑗
𝐿 = ∏𝑁
𝑗=1[[𝐹(𝑥𝑗 ; 𝑐𝑘 ; 𝜁𝑘 )] ] [[1 − 𝐹(𝑥𝑗 ; 𝑐𝑘 ; 𝜁𝑘 )] ] (3.8)

where, rj denotes the damage state of the bridge at damage state k under the ground motion
with PGA xj. It either takes magnitude equal to 0 or 1 depending on whether the bridge has
exceeded the defined threshold. The variability of the distribution indicates the scatternets of data.
To avoid the overlap of any two fragility curves, a single value of 𝜁𝑘 equal to 0.6 is used in this
study for all damage states. This quantity is suggested in FEMA (2013). As a result, the change in
median value serves as a measure for the variation in bridge fragility characteristics caused by
varying seismic loading. The fragility curves for the investigated components for three different
level of flood scenarios is illustrated in Fig. 3.11. The fragility median is inscribed within the
figure.

According to this figure, the only component of that could cause major damage and
collapse under seismic excitations for the three investigated flood scenarios is the pier. With an
increase in flood hazard level, there are noticeable changes in the seismic fragility of pier at the
major-damage and collapse states. This is solely due to an increase in scour depth, which increases
the flexibility of piers. With increasing flood hazard level, there is little to no change in seismic
vulnerability for all other cases, including minor and moderate damage of bridge piers. The same
scenario applies for the abutment and elastomeric bearing.

66
Figure 3.11 Component-level fragility curves of case-study bridge for four damage states at (a)
Pier - Flood scenario I, (b) Pier - Flood scenario II, (c) Pier -– Flood scenario III, (d) Elastomeric
bearing at all flood scenarios, (e) Abutment active deformation at minor damage state

3.7.2 System-level fragility curve

The system-level fragility curves for the case-study bridge are generated by combining the
component-level fragility curves into an overall system-level fragility curve. It is combined in such
a way that the failure probability of the system at particular damage state is the logical OR
operation of each failure probability of individual component (i.e., the failure of the system at any
damage state is considered when at least one of the components get fails). Fig. 3.12 shows fragility
curves of the bridge at four damage states under seismic hazard and for three flood scenarios.
Within a damage state (such as major damage or collapse state), curves at the extreme right and
left are the strongest and the weakest, respectively, as they provide the lowest and highest
probabilities of attaining a damage level under a given PGA. At lower damage levels of the bridge,

67
fragility curves corresponding to three flood scenarios overlap each other, indicating no variation
in bridge seismic vulnerability at minor and moderate damage states due to changing flood hazard
(Fig. 3.12a and Fig. 3.12b). However, notable variations in fragility characteristics of the bridge
are observed at major damage and collapse states (Fig. 3.12c and Fig. 3.12d). This is due to
expected intensive future flood events that would greatly impact the bridge at its higher seismic
damage levels.

(a) (b)

(b) (d)

Figure 3.12 Seismic fragility curve of the bridge for three flood scenarios at (a) minor damage,
(b) moderate damage, (c) major damage and (d) collapse state

68
3.8 Multi-hazard risk and resilience of case-study bridge

Risk and resilience-based analyses and design of bridges of highway transportation systems
account for the societal impact of bridge damage after a disaster. From several extreme events in
the past, it has been observed that beyond assessing bridge damage under any hazardous condition,
importance lays in estimating the damage consequences following the event and its negative
impact on the societal economy. By definitions, risk provides the measure of expected post-disaster
losses owing to bridge damage and resilience signifies its ability to revert back to the pre-damaged
functional level. Clearly, risk and resilience have their own significances in defining the impact of
bridge damage on societal backdrop, and should be estimated for an informed decision making on
bridge retrofit, rehabilitation and maintenance interventions towards climatic change adaptation.

For an effective adaptation to climatic change, risk-based design approach has now been
widely accepted by climate change experts (Bouwer 2019). The process deals with uncertainties
of the future climate projection model by considering the broader spectrum of damage and
associated losses (Caltrans 2018). In this section, the current study presents the calculation of
probable losses and simulation of post-disaster recovery process of the studied bridge. It is then
followed by subsequent discussions on the estimation of risk and resilience of the bridge under the
stated multi-hazard condition.

3.8.1 Direct and indirect losses

Losses incurred due to damage of a bridge following a disaster are classified in the
literature as direct and indirect losses. Chapter 2 discusses these two loss terms in detail. As
mentioned, direct loss represents the post-event cost required to repair damaged bridge
components. For a multi-hazard scenario ms, the post event bridge restoration cost CDL is estimated
by following the expression given in Zhou et al. (2010) and it is calculated based on expression
given in Eq. (2.8). Bridge replacement value Creb can be calculated as product of bridge deck area
and the average unit are replacement costs given by the California Department of Transportation
(Caltrans). Based on Caltrans construction statistics, average unit area replacement cost for RC
bridges is taken to be equal to $2960.1.6 /m2.

Besides direct loss, a damaged highway system suffers from indirect losses arising from
traffic delay and network downtime due to partial or full closure of damaged bridges in the

69
distressed network. This loss accumulates over time until all damaged bridges are fully restored
and the system is reopened to traffic at its full or pre-damaged functionality level. An exact
estimation of indirect loss is a challenging task as the process involves several factors that cannot
be defined with full certainty. The study here follows the model proposed by Stein et al. (1999) in
which traffic data and costs associated with detours are incorporated, and the numeric expression
was discussed in Eqn. (2.11) and Eqn. (2.12). Because the model is traffic-data based, values of
model parameters are obtained for the case-study bridge and tabulated in Table 3.3. These values
are kept unchanged for three different flood hazard scenarios. This assumption helps to study the
sole impact of climate change on bridge multi-hazard performance by eliminating the biasness of
several other analysis parameters.

3.8.2 Post-event recovery and restoration

Beyond loss assessment, an efficient management strategy for built infrastructures must
include a plan to be adopted for a quick post-event recovery and restoration. Post-event bridge
recovery is dependent on various money and human-driven factors; hence, no unique recovery
process exists in the reality that is generally applicable to any bridge damage scenario at any socio-
economic background. This is why a number of numerical models, continuous and/or discrete
functions, are proposed in the literature to simulate bridge post-disaster recovery. A review of
these existing models can be found in Chapter 2 (refer Section 2.8.3). This study adopts recovery
function proposed in Vishwanath and Banerjee (2019), which is discussed in detail in Section 2.8.3
of this dissertation. This recovery model for each damage state of the bridge is used to calculate
seismic resilience.

3.9 Multi-hazard risk of case-study bridge

Risk of the bridge is calculated at three levels of design earthquakes having exceedance
probabilities of 10%, 5%, and 2% in 50 years in the presence of three flood scenarios. These hazard
levels respectively correspond to minor, moderate and high levels of seismic hazard of the study
region, and henceforth referred to as level 1, 2 and 3. The reason behind selecting these three levels
of earthquakes is that these are generally been taken for probabilistic seismic hazard analysis and
used in most of the seismic design codes worldwide. Note that different other levels of seismic
hazard can as well be selected for this purpose.

70
From seismic hazard maps of the study region, PGA values of earthquakes corresponding
to seismic hazard level 1, 2, and 3 are found to be 0.30g, 0.38g, and 0.48g, respectively (refer Fig
3.9a). Based on these PGAs, probability of bridge damage at four damage states are obtained from
fragility curves (refer Fig. 3.12) and accompanying losses are estimated for various combinations
of seismic and flood hazards (a total of nine combinations). Fig. 3.13 presents these expected losses
in ascending order. The Fig. 3.13 shows high consequences from the multihazard scenario for low
probabilities of hazard occurrence, and thus emphasises the significance of calculating risk while
dealing with extreme events in association with climate change. For risk-based design and
decision-making purposes, risk curves are also generated for the bridge and plotted in Fig 3.14.
These curves provide annual exceedance probabilities of various level of expected losses from
bridge damage due to the multihazard scenario considered herein.

Table 3 3 Magnitude of Traffic Parameters Associated with Indirect Loss

Parameters Magnitude Source of Information

Operation Cost of Car (Cop,car) 0.4 ($/km) AASHTO (2003)

Operation Cost of Truck (Cop,truck) 0.6 ($/km) AASHTO (2003)

Average Daily Traffic (ADT) 80000 FHWA (2019)

Average Daily Truck Traffic Ratio (T) 26 % FHWA (2019)

Detour Length (Dl) 39 km Google Maps

Average Wage per hour (CAW) 18.26 ($/hr) U.S Dept. Of Labor (2018)

Average Car Occupancies (Ocar) 1.50 AASHTO (2003)

Average Truck Occupancies (Otruck) 1.05 AASHTO (2003)

Average Total Compensation per hour


28.91 ($/hr) U.S Dept. Of Labor (2018)
(CATC)

Time Value of Goods Transported in


3.81 ($/km) AASHTO (2003)
Cargo (Cgoods)

71
Parameters Magnitude Source of Information

Average Detour Speed (S) 45 km/hr Assumed

Length of Link (l) 2.414 km Google Maps

Average Speed of Car on Damaged Link


20 km/hr Assumed
(SD)

Average Speed of Car on Intact Link


90 km/hr Assumed
(SO)

Figure 3.13 Expected losses for various combinations of the multihazard event

72
Figure 3.14 Seismic risk curve of case study bridge for three scenarios of flood hazard

3.10 Multi-hazard resilience of case-study bridge

Resilience provides an integrated measure of structural capacity and adaptation in the


forms of “absorb” and “recover”, respectively. A systematic and comprehensive review done by
Banerjee et al. (2019) documents various methods and approaches generally taken by the bridge
engineering community to calculate resilience of bridges and highway transportation systems.
Here, the traditional approach of resilience assessment involving bridge vulnerability, loss and
recovery models is used for nine combinations of seismic hazard levels and flood cases. In the
calculation, functionality Q(t) of the bridge at a time instance t is expressed as given by Eq. (2.6)
and Eq. (2.7) in Chapter 2.

Fig. 3.15 shows resilience values of the bridge for nine combinations of multihazard cases.
This measure of bridge resilience has particular significance for the climate change scenario
because of the two-fold impact of climate change on the resilience assessment of the bridge. First,
global warming makes the bridge more vulnerable under seismic hazard (as can be seen from
fragility curves; refer Fig. 3.12) resulting in its higher seismic risk (refer Fig. 3.13). Second, due

73
to higher attaining damage level of the bridge under the same level of seismic hazard, recovery
takes longer time resulting in an additional rise of indirect losses.

Figure 3.15 Multihazard resilience of the bridge for the three scenarios of flood and seismic
hazard

3.11 Discussion of results – impact of climate change on risk and resilience of the bridge

Increased level of peak flow in future years due to climate change causes higher losses that
in turn results in larger risk of the bridge under the same multihazard level of earthquake and flood-
induced scour. Results presented in Fig. 3.13 and Fig. 3.14 for the case study bridge demonstrate
the fact. While Fig. 3.13 provides increments in multihazard risk due to climate stressors at specific
seismic hazard levels, risk curves in Fig. 3.13 portray a comprehensive assessment of annual
exceedance probabilities of expected losses in which the overall seismic hazard of the study region
is considered. For an example, there is an annual probability of 4×10-4 that losses due to the
multihazard will be equal to or more than $7.62 million for current climatic condition (i.e., flood
scenario I). For the same annual probability of being exceedance, the loss term becomes $8.65

74
million (13.5% rise) and $9.23 million (21.26% rise), respectively for flood scenario II and III.
Such a projected rise in multihazard risk of the bridge surely demands attention from the
engineering community for developing risk mitigation and adaptation plans to deal with the impact
of climate change in years to come.

The adverse impact of climate change is also observed on resilience of the bridge under the
same multihazard condition. As Fig. 3.15 shows, the change in flood scenario from I to III results
in 11% reduction of resilience (with respect to flood scenario I) at seismic hazard level 1. The
same is 14% at seismic hazard level 2. These reductions are not ignorable quantities; equivalent
reductions in resilience can be observed at flood scenario I (i.e., with no climate change projection)
when seismic hazard level at the bridge site changes from 1 to 2 (i.e., minor to moderate) and from
2 to 3 (i.e., moderate to high), respectively. Hence, the expected drop in resilience due to climate
change projection is indeed significant. This emphasises the additional threat of climate change on
long-term safety and serviceability of the bridge. Therefore, climate change adaptation becomes
an important aspect in this relation.

3.12 Climate change adaptation

The adverse effect of climate change on bridge systems can be controlled in two ways –
mitigation to reduce concentrations of atmospheric greenhouse gases and adaptation to minimize
the vulnerability and risk of bridges (NCHRP 2014). Adaptation is done with a target to observe
an immediate (i.e., near-term) effect that may exist over a long duration, whereas mitigation actions
are imposed early that make it easier to adapt effectively in the long-term. Hence, these two are
complementary activities in time scale (Yohe and Strzepek 2007). For constructing a new river
crossing bridge, adaptation for climate change is implemented at the design phase by assessing the
design flood to accommodate nonstationary climate trend into its design metrics. The same for an
existing bridge should be done during its ‘in-service’ phase by taking necessary repair and retrofit
measures in order to reduce its vulnerability in future.

The study here explores adaptation measures for the case study bridge to reduce its risk
from climate stressors. The preceding sections of this article demonstrated that climate change
induces higher depth of scour, and hence results in an enhanced multihazard vulnerability of the
bridge under earthquake and flood-induced scour. Thus, adaptation actions must include
appropriate scour countermeasures to retrofit the bridge for reducing its potential safety hazard

75
due to scour. Bridge retrofitting for scour is purely a site-specific assessment as it depends on a
variety of factors including the type of construction, river environment, maintenance history and
budget. For pier scour of existing bridges, Mondoro et al. (2018) listed in-service adaptation
methods that are currently in practice. Among these methods, Departments of Transportation in
the U.S. (USDOTs) suggest to install pile ripraps, which are basically piles of loose stones
armoured around bridge piers within the scour hole and above. Properly designed and constructed
ripraps are effective for long-term scour protection as well if periodically monitored and
maintained after each flood event (Lagasse et al. 2007). The same technique is adopted here as a
climate change adaptive measure for the case study bridge. The section below discusses its
application and potential benefits through a cost-benefit study performed over the remaining
service life of the bridge.

3.12.1 Application of rip-rap around bridge pier

Available riprap design methodologies for scour remediation provide a threshold size of
rocks that is expected to be stable under design floods. Such determination of critical riprap size
is important as it controls the shear failure under design flow. A number of empirical equations
are available in the literature for determining riprap size; however, there exists a large variation
among these equations due to limited real-life data based on which these equations were derived.
The fact is demonstrated in Lagasse et al. (2007) using existing riprap sizing equations for round-
nose piers. For more accurate assessment of riprap size under design flow, numerical analyses are
performed in past studies to simulate the field application of riprap around bridge piers. Such a
case study is presented in Suaznabar et al. (2017) in which an existing bridge in California with
oblong piers on pile foundations (attributes similar to those of the case study bridge) was
considered. That bridge was retrofitted by Caltrans in 2012 with a 1-ton average size rock riprap.
Suaznabar et al. (2017) developed 3D computational fluid dynamics (CFD) model of the bridge-
riprap system including the river bed, and analysed it under design flow to assess the stability of
applied riprap system during design flood events. Such a complex and computationally expensive
approach is not feasible to perform for every project aiming at riprap application. Hence, most of
the work done on bridge scour retrofit using ripraps depend on available riprap sizing formulas to
a major extent.

76
Richardson and Davis (1995) and Lagasse et al. (2007) suggest the use of Isbash equation
in determining the size of riprap for local scour protection. In this, the stone size is determined by
considering the critical velocity of stream flow near the boundary of bridge piers. The median
riprap size, dr50 (in m) is given by Eq. (3.10)

(𝐾 𝑉)2
1
𝑑𝑟50 = 0.692 (𝑆 −1)2𝑔 (3.10)
𝑠

where, K is the coefficient for pier shape (1.5 for round-nose pier and 1.7 for rectangular pier), Ss
is the specific gravity of stone used for riprap (2.65 for current case study), V is the mean approach
velocity (m/s), and g acceleration due to gravity (m/s2). To adapt the climate change-induced scour
risk to the furthermost, mean discharge from projected flood hazard corresponding to 2051-2099
is considered in the current study to determine the median riprap size for bridge piers. This is
obtained to be 0.88m for 100-year flood. The riprap is placed around bridge piers from the level
of footing to the river bed elevation as shown in Fig. 3.16. Since this design is done by considering
the highest projected design flood flow velocity over the life-cycle of the bridge, applied riprap is
expected to be stable under future flood events with frequencies higher than that of the design
flood. Further, it is expected that the bridge with ripraps behaves like as-built bridge (i.e., with no
scour at foundations) and the applied riprap prevents future erosion of soil material around bridge
piers. The sectional view of rip-rap retrofit is shown in Fig. 3.17.

3.12.2 Risk-based cost benefit of retrofitting

Benefit from riprap application can be described as the cost avoided from bridge post-
damage repair (i.e., direct cost) and network downtime (i.e., indirect cost). Essentially, this cost
avoidance happens as the retrofitted bridge performs better than before retrofitting under the same
multihazard condition. Following Chang et al. (2000), the expected annual benefit ( B ) due to
scour retrofit of the case study under regional earthquakes can be computed as given by Eq. (3.11)

0 𝑅) 0
𝐵 = ∑𝐸𝑒=1[(𝐶𝑆𝑒 − 𝐶𝑆𝑒 + 𝐶𝑅𝑒 𝑅 ]
− 𝐶𝑅𝑒 . 𝑝𝑒 (3.11)

77
Figure 3.16 Elevation view of bridge pier retrofitted with rip-rap

Figure 3.17 Section view of bridge pier retrofitted with rip-rap

78
0 0
where, CSe and CRe respectively represent indirect and direct losses under mth scenario earthquake
R R
when the bridge is not retrofitted. The same quantities after retrofitting are CSe and CRe ,
respectively. Pe represents the annual exceedance probability of eth seismic event. This calculation
should be performed for E number of scenario earthquakes that represent the regional seismicity.
These ground motions for the case study bridge are selected according to the regional seismic
hazard as provided by the US geological survey (USGS). After quantifying the expected monetary
benefit of scour retrofit ( B ), the total benefit (B) of retrofit for the remaining service life of the
bridge is calculated by considering certain discount rate on its present value. Hence, the total
benefit (B) is evaluated as expressed in Zhou et al. (2000) and it is given by Eq. (3.12)

𝐵 (1+𝐷𝑅 )𝑇 −1
𝐵 = ∑𝑇𝑡=1 (1+𝐷 )𝑡
= 𝐵𝐷 𝑇
(3.12)
𝑅 𝑅 (1+𝐷𝑅 )

where, DR and T represent, respectively, annual discount rate and remaining service life of the
retrofitted bridge. Generally, design service period of bridges is considered to be 75 years
(AASHTO 2017). As the case study bridge was constructed in 1972, cost-benefit analysis is first
done for T = 30 years and then repeated for a service life between 20 to 40 years.

Based on the total benefit, the cost effectiveness of scour retrofit in seismic enhancement
is evaluated through a Benefit Cost Ratio (BCR) that is calculated by dividing total benefit (B)
with total cost involved in scour retrofit (CR). Hence, any value of BCR higher than 1.0 signifies
profitable return from climate change adaptation. The life-cycle cost for a stone riprap includes
costs associated with construction materials, installation labour and periodic maintenance. A
typical cost of stone riprap in the central California was reported to be 80 to 120 USD per ton in
2003 (Griggs et al. 2005). The current study adopts an average value of 100 USD per ton of stone
riprap for the cost-benefit analysis. Table 3.4 enlists the outcome of this cost-benefit analysis
considering an average DR = 2.2% and T = 30 years. This average DR is calculated based on 10-
year statistics (from 2009 to 2018) of annual inflation rates of the U.S. (Coin News, 2015). Yearly
variation of the same yields a standard deviation (SD) of DR equal to 1.1%. Hence to estimate the
possible variation of mean BCR, the cost-benefit analysis is repeated by taking mean±SD values
of DR for T = 20 to 40 years. Obtained results, presented in Fig. 3.18, portray higher benefit from
retrofit as the retrofitted bridge provide service for several years after retrofitting and the annual
discount rate is minimal

79
Table 3.4 Cost-Benefit Analysis of Scour Retrofit for 30 Years Remaining Service Life of the Bridge and with Mean Discount Rate of
2.2%

Item No. Description Value

Cost of Installing Scour Countermeasure

1. Estimated Cost of Stone Riprap Raw Material (in Million USD) [76] 0.13

2. Estimated cost of scour monitoring device (in Million USD) [78] 0.0175

3. Cost for yearly maintenance (in USD) (assumed) 2500/year

4. Total retrofit cost (in Million USD) 0.22

Benefit-Cost Analysis

5. Expected annual benefit under regional seismicity ( B , in Million USD) 0.0325

6. Total benefit of scour retrofit (B, in Million USD) 0.71

9. Benefit Cost Ratio (BCR) of scour retrofitting [6]/[4] 3.18

80
Figure 3.18 Benefit cost ratio (BCR) for rip-rap application around bridge piers as a climate
change adaptive measure; dark line represents mean BCRs estimated based on an average
discount rate of 2.2% and the shaded area is for mean±SD

3.13 Closure

This chapter demonstrated the adverse effect of climate change on the multihazard
vulnerability, risk and resilience of a bridge over the San Joaquin River in California, USA. The
San Joaquin River is a primary river in California that meander through the Central Valley and
serve as an important source of water for the region. Due to climate change, the river is expected
to have higher discharge due to early runoff from melting snow. The design flood events in future
years are expected to intensify and result in higher scour depth around bridge piers. Owing to such
rises in design flood and associated pier scour, the case study bridge is expected to have higher
vulnerability and risk in future years under multiple hazard events involving earthquakes and
floods. If single hazard events (such as flood-induced scour alone) were to be considered, the
bridge might not have experienced significantly larger risk due to climate change projection;
however, the added risk because of climate change is not ignorable under the multihazard

81
condition. Consequently, resilience of the bridge under the same multihazard condition is observed
to have significant drop with time.

A climate change adaptation plan is implemented. The plan incorporated the application of
ripraps around bridge piers. It is designed such that the bridge can recover fully from existing scour
and no further scour occurs in future years under any flood event up to the level of design flood.
A cost-benefit analysis for the applied adaptation plan revealed its beneficial socio-economic
effect for a possible range of discount rates over a remaining service life of the bridge between 20
to 40 years. Overall, the study identifies climate change as a major additional threat to the
transportation infrastructure and signifies its critical impact on multihazard performance of river-
crossing bridges. The discussion presented here paves a way for future studies focusing on the
multihazard risk mitigation of bridges under climate change and the development of disaster
resilient bridge infrastructure systems.

82
This page is intentionally left blank

83
CHAPTER 4

IMPACT OF BRIDGE DESIGN PARAMETERS ON MULTI-HAZARD


PERFORMANCE-BASED DESIGN OF RIVER CROSSING BRIDGES

Previous chapter emphasized the importance of considering the interaction of earthquake


and flood-induced scour on the seismic performance of river-crossing bridges. It is observed that
bridge pier is the primary component that exhibits seismic vulnerability at higher damage states
and results in higher risk of bridges for different depths of scour. Consequently, this chapter
investigates the impact of different pier design parameters on the performance of river-crossing
bridges under the same multi-hazard condition. For this, the study adopted the performance-based
engineering framework with code-specified performance objectives for multihazard design of
river-crossing bridges. Thus, the study contributes to generate knowledge on specific roles of
bridge design parameters for multi-hazard design of riverine bridges and identifies key design
parameter(s) to meet specific risk criteria.

4.1 Multi-hazard condition at test bed region

Analyses are conducted by taking multihazard condition at the southeast part of Nepal as
test-bed scenario. Historically, Nepal has witnessed a number of major earthquakes. Starting from
the 1988 Udaypur earthquake to the 2015 Gorkha earthquake, the country has witnessed
tremendous economic losses arising from infrastructural damage, casualties, and disruption in
normal life flow (Gautam et al. 2019). In an article focusing on Himalayan earthquakes, Bilham
(2019) reported that a next big earthquake of magnitude 8 or greater is expected in this region
which may cause damage higher than the 2015 Gorkha earthquake. Moreover, various intensity
flood events (i.e., extreme and recurrent) caused significant economic losses in the Koshi river
basin located in the southeast part of Nepal. The field reconnaissance survey of bridges affected
by the 2017 flood in Nepal indicates that settlement and scouring around bridge piers were
commonly observed for bridges that experienced floods. Hence, pier scour along with the strong
earthquake will put bridges in Nepal at risk due to multiple hazards during their service lives

84
(Adhikari et al. 2019). This section discusses the selection of seismic and flood hazard models for
this test-bed region.

4.1.1 Regional seismicity and selection of ground motions

Selection of earthquake ground motions plays an essential role in developing fragility


curves through nonlinear response-history analysis. Fragility curves are affected by the presence
of randomness among seismic records, and thus ground motions that are recorded or simulated at
a region of interest should be used to conduct vulnerability analysis for that region. Recorded
ground motions in the southeast part of Nepal are not enough to utilize for seismic vulnerability
analysis of bridges. Thus, additional ground motions representing Nepal’s seismic potential are
explored in the PEER ground motion database (PEER, 2023). Records from the database were
chosen from around the world with parameters that represent the same seismic characteristics and
fault mechanism as in the Nepal region. The selection criteria include moment magnitude (Mw)
ranging between 5 to 8. These were chosen based on earthquakes that occurred in Nepal since
1911, and the region’s fault mechanism consists primarily of reverse, reverse-normal, and reverse-
strike slip (Shrestha, 2014). The site condition at the region consists of soft soil deposits, typical
of NEHRP site class D classification; thus, shear wave velocity was chosen between 180 and 360
m/s (Goda et al. 2015). In addition to the selected sets of ground motions, recorded ground motions
from the 2015 Gorkha earthquake, obtained from a recently published dataset (Shigefuji et al.
2022), are included in the ground motion dataset.

To select region-specific ground motions, seismic spectral acceleration developed by


Chaulagain et al. (2015) is utilized. Based on the developed curves, response spectra (5% damping)
with a specified probability of exceedance (say, 10%, 5%, and 2% in 50 years) are used as target
spectra for collecting ground motions from the PEER database. In this process, 60 sets of ground
motions from the PEER database are obtained following the ground motion selection criteria.
Average response spectra of each suite of ground motions are shown in Fig. 4.1. Tables C1 to
Table C3 in Appendix C of this dissertation list these ground motions for each seismic hazard level
and corresponding peak ground accelerations. Additional 16 sets of ground motions, recorded in
the southern part of Nepal during the 2015 Gorkha earthquake, are considered in the dataset (also
included Table C4 and Table C5 in Appendix C of this dissertation). Thus, a total of 76 ground
motions are used for seismic vulnerability assessment of bridges studied here.

85
Figure 4.1 Generated earthquake ground motions based on the seismic hazard at the testbed
region

4.1.2 Characteristics of 100-year flood with climate change projection and scour depth

The Kochi River is one of the major rivers in the southeast part of Nepal. Based on the
historic flow in the river, Devkota and Gyawali (2015) reported the discharge of 100-year flood in
this river is 6470 m3/s for the observation period 1997-2006. This literature also considered two
Regional Climate Models (RCMs) to estimate projected flood discharge in the river due to the
global warming and change in climate. According to the climate change projection in the Kochi
River basin area, 100-year flood discharge for projection period 2041-2050 was obtained to be
7625 m3/s when one of the RCMs was used. This flood characteristics (with climate change
projection) is used in the current study to estimate the depth of scour at around bridge piers. For
the adopted flood discharge, scour depth around bridge foundations is estimated using the
formulation given in Eq. (3.7).

4.2 Description of representative bridge and design parameters

The case-study bridge is assumed to be located in Nepal. From the database of bridge
management system in Nepal (BMS, 2022), it is observed that three-span bridges consume
approximately 27% of river-crossing Nepali bridges having more than one span. Moreover, the
average length of these three-span bridges is 51 m with a standard deviation of 36 m. Following
that, the current study developed a representative three-span bridge with an overall length of 86 m
that approximately becomes equal to mean+1SD of the observed statistics on overall length of
three-span river-crossing bridges in Nepal. Two exterior spans of the bridge are 25 m long while

86
the interior span is 36 m long. Basic geometry, cross-sectional details, and major nonlinear
components of the bridge are shown in Fig. 4.2. The basic bridge model is designed in accordance
with AASHTO (2017) LRFD bridge design specifications. Bridge girder is monolithically
connected to pier bents and are supported on elastomeric bearings and seat-type abutments at two
ends. At each bent, there are two piers of 12 m height and 1.2 m diameter. These are supported on
pile groups each having 8 piles each of 0.9 m diameter and a pile cap.

Figure 4.2 Geometry, cross-sectional details, and major nonlinear components of the base model

In performance-based analysis and design of bridges, aspect ratio for piers (L/D, where L
and D are effective height and diameter of piers) and longitudinal reinforcement ratio (ρL) in piers
are two key bridge design parameters. Lehman and Moehle (2000) performed experimental study
of RC piers by varying L/D between 4 to 10 and ρL between 0.75% to 3.0% to gauge their
influences on response and failure of bridge piers in the context of performance-based seismic
design of bridges. Following that, the current study considered these two design parameters for
bridge piers and explore their confronting roles on multihazard performance of river-crossing
bridges. As these parameters can vary in the field, 10% variation for each is considered here as
presented in Table 4.1. For each structural condition, two bridge models with and without scour
are developed. Thus, a total of 14 bridge models are generated based on the basic model (or, base

87
model, BM) in Table 4.1. For all cases between Br. 1 to 10, pier length is kept constant at 12 m.
Scour depths at arounds bridge piers are estimated following the relation presented in Eq. [3.7].
Obviously, higher pier diameter results in higher scour depth and vis-a-versa. That is why Br. 4
has higher and Br. 6 has lower scour depths compared to that in Br. 2 under the same flood event.
The longitudinal reinforcement ratio (i.e., 𝜌𝐿 ) is also varied by 10% from its value used in the BM.
Note that when one of these two ratios (i.e., L/D and 𝜌𝐿 ) is varied, the other one is kept at its value
used in the BM. Lastly, two values of differential ground elevation (or, pier height) between bents
are considered, although the scour depth at all piers in Br. 12 and 14 are kept same as those for Br.
2. Through the imposed variations amongst the three parameters (as shown in Table 4.1), their
individual impacts on the multihazard vulnerability and risk of bridges are determined. The
description for developing the FEM of a representative bridge in OpenSees is discussed in Chapter
3 (refer to section 3.3), and the same modelling strategy is used to develop the FEM in this Chapter.

Table 4.1 list the estimated scour depth along with other specifications of the bridge models
under consideration. Note that time-dependent scour is not studied in the current article. The
estimated scour depths are due to local scour arising from one-time flood event in the river.

4.3 Impact of bridge design parameter on moment-curvature relation of pier

The moment-curvature (M-ϕ) relationship can be used to calculate the maximum bending
moment and deformation of a pier under different loadings before it fails. In performance-based
design, engineers can design the structure to meet specific performance criteria, such as limiting
the curvature of the pier to prevent excessive deformation or damage, by utilizing the M-ϕ
relationship into account. For bridges with different pier cross-sectional properties (i.e., for Br. 1,
3, 5, 7, and 9), M-ϕ response of pier cross-sections are obtained until the ultimate point when
extreme fibers of confined concrete section reach the ultimate strain level. These M-ϕ plots, actual
and idealized, are shown in Fig. 4.3a and Fig. 4.3b in which solid lines represent actual relations
and corresponding idealized relations are presented with dashed lines. Note that, scour depth and
differential ground elevation between piers do not impact the cross-sectional property of piers.
Thus, the M-ϕ response of Br. 1 essentially represents the same for Br. 2, 11, 12, 13, and 14.

88
Table 4.1 Specifications of the bridge models under consideration

Longitudinal Differential
Pier reinforcement
Bridge ID Scour ground
Description diameter ratio (𝝆𝑳 )∗
# depth (m) elevation
(m)
(%) between bents

Base model
Br. 1 1.20 - 1.8 -
(BM)
Base model -
Br. 2 with scour 1.20 3.73 1.8
(BMS)

BM with 10% -
Br. 3 1.32 - 1.8
higher diameter

BMS with 10% -


Br. 4 1.32 3.97 1.8
higher diameter

BM with 10% -
Br. 5 1.08 - 1.8
lower diameter

BMS with 10% -


Br. 6 1.08 3.48 1.8
lower diameter

BM with 10%
higher -
Br. 7 1.20 - 2.0
reinforcement
ratio
BMS with 10%
higher -
Br. 8 1.20 3.73 2.0
reinforcement
ratio
BM with 10%
lower
Br. 9 1.20 - 1.6 -
reinforcement
ratio

89
Longitudinal Differential
Pier reinforcement
Bridge ID Scour ground
Description diameter ratio (𝝆𝑳 )∗
# depth (m) elevation
(m)
(%) between bents

BMS with 10%


lower -
Br. 10 1.20 3.73 1.6
reinforcement
ratio

BMS with
Br. 11 differential pier 1.20 3.73 1.8 0.5
height – 1

BMS with
Br. 12 differential pier 1.20 3.73 1.8 0.5
height – 1

BM with
Br. 13 differential pier 1.20 - 1.8 0.8
height – 2

BMS with
Br. 14 differential pier 1.20 3.73 1.8 0.8
height – 2

* of gross column cross-sectional area (Ag) of BM

For the same reason, M-ϕ relations for Br. 3 and 4 are the same, Br. 5 and 6 are the same,
Br. 7 and 8 are the same, and Br. 9 and 10 are the same. As Fig. 4.3a shows, the linear elastic
portions of the plots have almost the same flexural stiffness; however, yield capacity changes
depending on sectional characteristics. The moment capacity of Br. 3 piers is comparably higher
than those in the base model (Br. 1) due to 10% pier diameter of Br. 3 piers. Similarly, Br. 5 piers
with 10% lower diameter than piers in the base model have lower moment capacity than the base
model. The change in yield curvature between these three bridges in Fig. 4.3a is marginal, though
a slight variation exists among the ultimate curvature values. Higher the pier diameter, lower is
the ultimate curvature and so is the curvature ductility. This is reasonable as lateral confinement
of a pier is expected to decrease as its diameter increases (without any change in other sectional

90
properties), and this results in lower curvature ductility. These observed trends in M-ϕ response of
bridges with varying pier diameter are also in accordance with the formulation given by Priestly
et al. (1995). The effect of change in longitudinal reinforcement ratio (i.e., 𝜌𝐿 ) on M-ϕ relations is
shown in Fig. 4.3b for Br. 1, 7, and 9. As these bridges have the same pier diameter, yield moment
capacity is the maximum for Br. 7 that has 10% higher 𝜌𝐿 in piers than that in the base model, Br.
1. For the same reason, Br. 9 (with 10% lower 𝜌𝐿 in piers than Br. 1) has the lowest value of yield
moment capacity compared to other two bridges. However, the imposed variation in 𝜌𝐿 has
marginal impact on the yield and ultimate curvature values. Thus, it can be stated that 10% change
in 𝜌𝐿 has no practical impact on curvature ductility of pier cross-sections.

Figure 4.3 Moment-curvature relations of bridge piers for varied (a) diameter and (b)
longitudinal reinforcement ratio

4.4 Nonlinear time-history analyses of bridges and seismic fragility curves

4.4.1 Limit state of various bridge component

Modern design codes have established and specified performance objectives for bridge
components with design level of earthquakes. For example, Canadian Highway Bridge Design
Codes (CHBDC, 2014) specify a design level earthquake with a return period of 475 years for
lower-level design and 2475 years for upper-level design, whereas bridges designed according to
the American Association of State Highways Officials (AASHTO 2017) specifications requires

91
life safety (1000-year return period) as the performance requirement (Zhang and Alam, 2019).
Therefore, it is important to develop numerical correspondence between seismic response of
bridges for given seismic hazards with their code-specified target performance levels. Based on
FEMA (1999) specifications, seismic damage of bridges is classified in this study similar to
damage states defined in Chapter 3 in five different seismic damage states (i.e., no, minor,
moderate, major damage and collapse/complete damage) according to the physical damage that a
bridge gets during seismic excitations. For identifying these damage states from bridge seismic
response obtained through numerical simulations, literature-specified threshold limits for various
bridge components in each of these damage states are used. Thus, curvature ductility (or sectional
ductility) of bridge piers, translational displacements of abutments and bearings are used to
measure seismic response of bridges (referred to as engineering demand parameters, or EDPs)
based on which seismic damage of bridge components at various performance levels are defined.
These damage states and corresponding threshold values are summarized in Chapter 2 (refer Table
2.3 and Table 2.4).

4.5 Variation of component response under a given seismic demand

For the selected set of 76 ground motions, nonlinear time-history analyses of all 14 bridges
are performed. Response quantities for each EDP are recorded, which show significant variations
between response values obtained for difference bridges under the same earthquake. For an
example, Fig. 4.4 shows the maximum curvature ductility values obtained for bridge piers with
and without scour when subjected to five strongest ground motions from the ground motion suite.
In this, Fig. 4.4a, Fig. 4.4b, and Fig. 4.4c compare values for bridges with, respectively, varying
pier diameter, varying 𝜌𝐿 , and varying (equal and differential) ground elevation at piers bases. In
each figure, two sets of vertical bar charts are shown for each ground motion – the first set is for
bridges without scour and the adjacent second set is for corresponding scoured bridges. As
expected, bridges with foundation scour are observed to have higher ductility demand than that for
respective bridges without scour (e.g., Br. 2 vs. Br. 1) due to increased flexibility resulted from
higher exposure length of scoured piers. Comparing results for bridges with different pier
diameters (from Fig. 4.4a), average decrease and increase in ductility demand of 10% and 6% are
observed for bridges, respectively, with 10% higher (for Br. 3 and 4) and 10% lower (for Br. 5 and
6) diameter piers than those obtained for base models. Similar variations in ductility demand for

92
bridges with mean ± 10% variation in 𝜌𝐿 , are comparatively less than that in Fig. 4.4a. As Fig.
4.4b shows, 8.1% decrease and 13.4% increase in ductility demand are observed, respectively, for
bridges with 10% higher (for Br. 7 and 8) and 10% lower (for Br. 9 and 10) longitudinal
reinforcement ratio (𝜌𝐿 ) in piers when compared to the values obtained from the base models.
These results are consistent with the moment curvature relation observed in Fig. 4.3, and Fig. 4.4c
does not show any variation in ductility demand of piers due to differential ground elevation at
pier bases. The variations observed in this figure between two sets of vertical bar charts for each
ground motion is solely due to the presence of scour at around bridge piers of Br. 2, 12, and 14.
Similar observation is made based on the longitudinal translation of elastomeric bearings.

Figure 4.4 Curvature ductility demand of bridge piers under seven strong ground motions;
bridges with (a) varied pier diameters, (b) varied longitudinal reinforcement ratio, and (c) same
or different riverbed elevation at pier bents.

4.6 Seismic vulnerability of investigated bridges

Fig. 4.5 and Fig. 4.6 show the fragility curves of bridges without and with scour,
respectively, at four damage states which are developed based on discussion presented in previous
chapters. Corresponding median fragility parameters are inscribed within the figures. For all cases,
the standard deviation is set to 0.6 (following FEMA (2003)) at all damage states to avoid
intersection between curves. It also helps to compare the curves based on their median values. As
these curves show, seismic vulnerability of non-scoured bridges does not vary significantly due to
the variations in pier design parameters and differential ground elevation at pier bases (Fig. 4.5).
However, observed variations are significant for scoured bridges at the major damage and collapse

93
states (Fig. 4.6). Differential elevations of 0.5m and 0.8m at pier bases do not produce any variation
in seismic and multihazard vulnerability of bridges. This observation goes as par with that from
Fig. 4.4c. For remaining 10 bridges, notable variations are observed at all damage states. Thus,
these 10 bridges are used in the next section to discuss specific roles of the design parameters on
seismic and multihazard performance of bridges.

Figure 4.5 Seismic fragility curves of bridges without scour at (a) minor damage, (b) moderate
damage, (c) major damage, and (d) collapse state

94
Figure 4.6 Seismic fragility curves of bridges with scour at (a) minor damage, (b) moderate
damage, (c) major damage, and (d) collapse state

4.6.1 Role of design parameters on bridge fragility curves

Role of pier design parameters on bridge seismic vulnerability is measured by comparing


fragility curves obtained for first 10 bridges. Fig. 4.7a, Fig. 4.7b, and Fig. 4.7c illustrate percent
changes in median fragility parameters, respectively, for non-scoured bridges when compared to
that of BM, for scoured bridges when compared to that of BMS, and scoured bridges when
compared to that of BM. Results in these figures demonstrate specific impacts of (i) design

95
parameters for non-scoured and scoured bridges (Fig.4.7a and Fig.4.7b), (ii) multihazard load case
(for Br. 2; Fig.4.7c), and (iii) the combined effect of multihazard and design parameters (Figure
8c, other than Br. 2) on bridge vulnerability. Note that when fragility parameters from scoured
bridges are compared within the group (Fig.4.7b), the result signifies the change in seismic (not
multihazard) vulnerability of bridges with exposed foundations (owing to scour). Central lines of
these plots stand for base lines representing values from the base models; positive and negative
changes from central lines express, respectively, decrease and increase in seismic vulnerability of
bridges with respect to the respective base model. As expected, the presence of flood-induced
scour results in higher seismic vulnerability of the base model (Br. 2 in Fig.4.7c).

As can be seen from Fig. 4.7a, 10% increase in pier diameter resulted in 6.7%, 5.9%, and
10.4% improvement in seismic fragility of Br. 3 at the minor damage, moderate damage, and
collapse states, respectively. Similar improvements for scoured bridges (i.e., Br. 4) are 5.6%, 3%,
10.6%, and 10.4%, respectively, in minor, moderate, major damage, and collapse states (Fig. 4.7b).
These improving trends of bridge seismic vulnerability indicate that increased pier diameter helps
enhancing seismic performance of bridges when scoured and non-scoured bridges are compared
within their respective groups. Decrease in pier diameter, on the other hand, results in enhancing
seismic vulnerability of bridges. In this case, higher degradation is observed for Br. 6, primarily at
the major damage (which equals to 26%) and collapse (which equals to 12.3%) states (Fig. 4.7b)
than that for Br. 5 in Fig. 4.7a. These observations are in accordance with the trends observed in
Fig. 4.4a for bridge response values under strong motions. Thus, it is understood that lowering the
diameter without changing other design parameters of piers results in lower moment carrying
capacity of pier cross-sections (Fig. 4.3a) and enhanced seismic vulnerability of bridges without
and with scour. In case of multihazard scenario (Fig. 4.7c), the impact of ±10% variation in pier
diameter on bridge fragility at minor and moderate damage states does not become eminent as
multihazard load case makes much higher impact than this design parameter. However, changes
are noticeable at the major damage and collapse states. In these higher damage states, larger pier
diameter helps to improve multihazard performance of the bridges, while lower diameter results
in enhance multihazard vulnerability of bridges (Br. 4 and 6 in Fig. 4.7c).

A 10% increase in the longitudinal reinforcement ratio 𝜌𝐿 of piers found to have no impact
on seismic vulnerability of bridges, except for the minor damage state at which 3 – 5% positive
changes in fragility are observed (Br. 7 and 8 in Fig. 4.7a and Fig. 4.7b). Similar marginal impacts,
96
but negative, are observed for Br. 9 when 𝜌𝐿 is lowered by 10% (Fig. 4.7a); however, negative
impacts are substantial at major damage (22%) and collapse (11%) states for Br. 10 (Fig. 4.7b).
As also observed earlier (in Fig. 4.4b), Br. 10 always produces higher ductility demand than other
bridges under strong motions; this trend is followed in fragility curves as well. Thus, it can be
inferred that lowering 𝜌𝐿 in bridge piers may become more detrimental for scoured bridges than
the without scour case. The observations from the multihazard scenario (for Br. 10 in Fig. 4.7c)
also confirm this. It is interesting to note from the above discussion that lowering design parameter
values (either the aspect or the longitudinal reinforcement ratio; Br. 6 and 10) enhances
multihazard vulnerability of river crossing bridges at the major damage and collapse states. Thus,
lowering design parameters may produce the worst impact on bridge vulnerability when
multihazard condition is considered.

Figure 4.7 Percent variation in fragility median at various damage states due to (a) variations in
design parameters in non-scoured bridges, (b) variations in design parameters in scoured bridges,
and (c) combined effect of scour and variations in design parameter

4.6.2 Role of design parameters on multi-hazard risk of investigated bridges

As observed from the preceding section, multihazard vulnerability of bridges gets


influenced by variations in diameter and 𝜌𝐿 of bridge piers. Therefore, it is imperative to explore
the extent of impacts these two design variables produce on multihazard risk of bridges. Moreover,
estimated risk will provide practicing engineers a quantitative basis for risk-based design of

97
bridges following PBE framework. Estimated risk provides a combined measure of bridge
vulnerability and the post-event consequences. In the study here, consequences are measured in
terms of monetary losses owing to post-event repair of damaged bridge components. In reality,
post-disaster economic losses also arise from indirect consequences such as traffic delay, network
downtime, and opportunity loss. However, bridge owners are mostly concerned about direct
consequences, and thus, traffic-data based indirect losses are not considered in the risk assessment
framework here.

Post-event repair cost (CRm) as a percentage of reconstruction cost of a bridge for with a
seismic event are expressed as risk curves and it is quantified based on the numeric expression
discussed in Eq. (2.8). Fig. 4.8 shows seismic risk curves of considered bridges (Br. 1-10). To
develop these curves, seismic hazard curve for the Katmandu region of Nepal is used as obtained
from Stevens et al. (2018). These risk curves represent the annual exceedance probabilities of
different values of bridge restoration cost that are expressed in terms of percentage of
reconstruction cost. For an example, there is 0.14% annual probability that repair to reconstruction
cost ratio for Br. 1 will be 28.86% or higher due to future seismic activity in the region. Thus, these
curves can also be used to compare seismic risk associate with multiple bridges located in the same
region. Higher the cost value for a given exceedance probability or higher exceedance probability
for a given repair cost level represents higher risk to the bridge. Fig. 4.8a and Fig. 4.8b portray
seismic risk of various bridges without scour as compared to that for the BM (i.e., Br. 1). Thus,
these plots show the sole impact of bridge design parameters on seismic risk of non-scoured
bridges. In Fig. 4.8c and Fig. 4.8d, the seismic risk is compared between bridges with scour and
the BM (i.e., Br. 1). These plots, therefore, signify the variation in seismic risk of scoured bridges
as well as the multihazard risk due to the variations in design parameters. While higher seismic
risk of Br. 6 compared to that of Br. 2 is the sole impact of lowering pier diameter of Br. 6, the
same when compared with Br. 1 can be attributed to the combined impact of lower pier diameter
and the multihazard load case (Fig. 4.8c). Similarly, the higher risk for Br. 10 compared to Br. 2
and Br. 1 signify, respectively, the sole impact of lowering 𝜌𝐿 in the pier and the combined impact
of multihazard and lower 𝜌𝐿 (Fig. 4.8d). With 10% increase in 𝜌𝐿 , seismic risk of scoured bridges
does not change (Br. 2 vs. Br. 8), whereas multihazard risk still increases (Br. 1 vs. Br. 8). For
non-scoured bridges, ±10% variations in these design parameters do not impose any significant
change in seismic risk (Fig. 4.8a and Fig. 4.8b), except for Br. 3 which is observed to have lower

98
risk than Br. 1 when pier diameter is increased. Higher pier diameter in scoured bridges (Br. 4)
also helps to reduce seismic risk of scoured bridges (Fig. 4.8c). All these observations demonstrate
that pier aspect ratio and 𝜌𝐿 are both key parameters for multihazard design of riverine bridges.

Furthermore, seismic risk of bridges is evaluated for three design level ground motions
following the performance-based design approach – service level earthquakes (SLE) with 50%
probability of exceedance in 50 years (i.e., 75-year return period), design level earthquake (DLE)
with 10% probability of exceedance in 50 years (i.e., 475-year return period), and maximum
credible earthquakes (MCE) with 2% probability of exceedance in 50 years (i.e., 2475-year return
period). For these three design levels of seismic ground motions, corresponding performance
levels for bridges – fully operational, operational, and life safety – are formulated based on FEMA
(2006) recommendations for bridges. Each of these performance levels is defined by a specific
percentage of repair cost, which is referred to as acceptable seismic risk. This study adopts the
acceptable seismic risk corresponding to each performance level of investigated bridges from
FEMA 451B (2007). Following a seismic event, a bridge is said to be fully operational when the
required repair cost is minimal (say, less than 5%), and the bridge is expected to be fully functional
for all types of traffic after the seismic event. A bridge is said to have operational performance
level after a seismic event when its required repair cost is greater than 5% but less than 15%. This
performance level signifies that the bridge gets repairable damage due to seismic activity and post-
event repair may require no/partial closure of the bridge. Residual strength of a bridge at this
performance level allows emergency vehicles on the bridge. Finally, a bridge has life safety
performance level when it suffers from substantial seismic damage that would require full closure
of the bridge from traffic due to significant repair or replacement of one or more bridge
components. Therefore, bridge repair cost at this performance level may exceed 15% but limit to
30% of its reconstruction cost. If the loss exceeds 30% of its reconstruction cost, the bridge may
be regarded to have reached collapse prevention performance level and any post-event repair may
not be practical.

Table 4.2 and Table 4.3 shows the percentage variation in cost incurred or avoided of non-
scoured and scoured bridges relative to the acceptable seismic risk for the three design seismic
levels. Here, positive values measure percent lower and negative values signify percent higher risk
compared to the corresponding acceptable limits. It is observed that seismic risk of all non-scoured
bridges is within the acceptable limit under SLE, but crosses acceptable limits for DLE and MCE.
99
For the scoured bridge, however, there is no bridge for which the acceptable risk criterion for any
performance level is met. Thus, in order to make seismic and multihazard risk of the investigated
bridges within the acceptable limits, appropriate changes in bridge design need to be performed.
Observed trends on impact of individual design parameter on ductility demand, seismic
vulnerability, seismic and multihazard risk of bridges guide to make appropriate decisions in this
regard.

Figure 4.8 Variation in multihazard risk with respect to that of the base model for bridges with
(a) varied pier diameter, (b) varied pier diameter in the presence of scour, (c) varied longitudinal
reinforcement ratio, and (d) varied longitudinal reinforcement ratio

100
Table 4.2 Percentage change in the repair cost of non-scoured bridge models relative to acceptable seismic risk under three design
seismic events

Seismic hazard level Br. 1 Br. 3 Br. 5 Br. 7 Br. 9 Br. 11 Br. 13

SLE 16.0 % 26.8 % 13.6 % 17.1 % 12.3 % 16.0 % 16.0 %

DLE -44.2 % -26.4 % -44.4 % -44.1 % -46.0 % -44.2 % -44.2 %

MCE -37.2 % -23.5 % -37.3 % -37.2 % -37.9 % -37.2 % -37.2 %

Table 4.3 Percentage change in the repair cost of scoured bridge models relative to acceptable seismic risk under three design seismic
events

Seismic hazard level Br. 2 Br. 4 Br. 6 Br. 8 Br. 10 Br. 12 Br. 14

SLE -86.0 % -71.2 % -129.2 % -85.1 % -121.1 % -86.0% -86.0 %

DLE -78.4 % -53.9 % -134.4 % -78.3 % -125.5 % -78.4 % -78.4 %

MCE -50.5 % -32.1 % -85.3 % -50.4 % -80.4 % -50.5 % -50.5 %

101
4.7 Closure

This chapter discusses the role of three key design parameters for bridge piers such as
aspect ratio, longitudinal reinforcement ratio, and differential ground elevation at pier bases was
explored on multihazard performance of riverine bridges. Appropriate variations of these
parameters are considered to estimate the extents of their impacts on bridge seismic vulnerability
and risk curves. A total of fourteen bridge models are developed with combinations of these three
design parameters in the presence and absence of bridge sour at foundations resulted from 100-
year flood event. Characteristics of the flood event was obtained from a past research that
considered historical annual peak flow discharge data and future flood projection based on a
regional climate change model. Results obtained from finite element analyses of the investigated
bridges under regional seismic hazard is expressed in the form of fragility and risk curves, and the
performance-based engineering framework is adopted to judge if estimated risk of bridges satisfies
code-specified performance objectives for different seismic hazard levels. Comparing obtained
risk values with code-specifies performance levels showed that risk of all investigated bridges is
within the acceptable limit for service level earthquakes, however, crosses acceptable risk limits
under higher level earthquakes. This indicates the need for revising design of bridges in order to
make them perform at the targeted performance level under specified hazard condition. Knowing
individual impact of different design parameters on seismic and multihazard performance of
bridges, as done here, facilitates informed decision-making on that aspect.

102
This page is intentionally left blank

103
CHAPTER 5

IMPACT OF COMBINED CORROSION-FATIGUE ON SEISMIC


VULNERABILITY AND RESILIENCE OF HIGHWAY BRIDGES

Previous chapters of this dissertation examined the seismic performance of river-crossing


bridges by considering the interaction between earthquakes and floods. In addition to this multi-
hazard condition, seismic performance of urban bridges is affected by the combined interaction of
corrosion and traffic-induced fatigue (as discussed in Chapter 1). This chapter thus focuses on the
development of a numerical framework to estimate seismic resilience of RC bridges subjected to
corrosion-fatigue degradation throughout their service lives. Upon development, the framework is
demonstrated to RC bridges having two and three spans, and the analyses provide insights into the
seismic vulnerability and resilience of these bridges over their service life. Overall, this research
indicates the role of corrosion-fatigue degradation and emphasizes its detrimental impact on the
life-cycle seismic resilience of RC bridges.

5.1 Overview of numerical framework for estimating seismic resilience of bridges


incorporating corrosion-fatigue degradation mechanism

A quantitative framework is developed here to assess the influence of corrosion-fatigue


degradation on life-cycle seismic performance and resilience of RC bridges. The proposed
framework, shown in Fig. 5.1, integrates various modules for analysing stochastic traffic load,
vehicle-bridge interaction (VBI), fatigue stresses, and material degradation owing to corrosion-
fatigue phenomenon. It is applied to RC T-girder bridges having girders and piers subjected to,
respectively, corrosion-fatigue and corrosion (alone) degradations. In the process, random gross
vehicle weights (GVWs) on the bridges are generated based on the information acquired from
recorded weigh-in-motion (WIM) measurements. Appropriate traffic growth models are applied
to account for the time-evolving nature of traffic cycles over bridge service life. Furthermore, time-
variant reductions in rebar area and elastic modulus of concrete in bridge girders are estimated
such that the effect of corrosion-fatigue is realistically captured. Obtained information on material

104
Figure 5.1 Flowchart for calculating material degradation of aging RC girders due to corrosion-
fatigue phenomenon

degradation is used in nonlinear time history analyses of the bridges at various degraded
conditions to obtain their time-dependent seismic fragility curves and resilience. The chapter first
discusses the methodology for modelling corrosion-fatigue degradation for RC T-girders, followed
by its application to multi-span bridges to calculate life-cycle seismic resilience.

5.2 Modelling corrosion-fatigue degradation on RC girders

5.2.1 RC bridge girder and traffic data

Multi-span simply supported T-girder RC bridges with span length equal to 20m are chosen
as case study bridges. Cross-sectional dimensions at mid-span of bridge decks and that of one T-

105
girder are adopted from Yan et al. (2017) and shown in Fig. 5.2. Further details on these bridges
and various bridge components are discussed later in this chapter. Compressive strength of
concrete and elastic modulus of reinforcing steel are taken as 40MPa and 195GPa, respectively.
Yan et al. (2017) reported that the longitudinal rebars located at the bottom of girders are fatigue-
critical rebars. Following that, fatigue-critical rebars for the considered bridge girder are estimated
to locate at a depth of 1.21m from the centroid of T-girder section (shown in Fig. 5.2).

Figure 5.2 (a) Dimension of the bridge deck at mid-span cross-section and (b) dimension of a
girder

For traffic analysis, recorded WIM data over a time span of several years is desirable to
observe the past traffic growth with time, and accordingly, time-evolving traffic volume for future
years can be projected. However, such data is not always available for every corridor of major
highways. To overcome this issue, the current study assumes that case study bridges experience
traffic volume similar to that observed for the Interstate LA-710 in Los Angeles, CA over a period

106
of 1995 to 2001 (Lu et al. 2002). Collected WIM data comprises of various types of truck traffic,
based on which stochastic traffic load on bridge girders is simulated and used for VBI analysis.

5.2.2 Sampling of vehicles and stochastic GVWs

The stochastic traffic load analysis considers a variety of random variables including
vehicle type and weight; among these, randomness in vehicle axle weight influences fatigue
stresses of bridge girders in the most significant way (Lu et al. 2002). Collected traffic data from
the Interstate LA-710 comprises of various types of trucks and their varying axle weights (Lu et
al. 2002). Table 5.1 provides descriptions and occupant percentages of observed truck traffic for
the highway corridor. As observed, five-axle Truck Type – 9 having steering axle followed by two
sets of tandem axles dominates the composition with 56.4%. It is followed by two-axle (steering
and single) Truck Type – 5 with 16.8% occupancy and three-axle (steering and tandem) Truck
Type – 6 with 14.2% occupancy. Remaining composition of the observed truck traffic was a
mixture of other trucks. These proportions of various truck types are used to generate random
samples of GVW.

Table 5.1 Truck types and its percentage occupancy in observed traffic volume

Vehicle Type
Illustration of its Axle (Short description)
(Occupancy)

Truck Type-9
(56.4%)
Five-axle trucks with single trailer (tractor with tandem axle and single semi-trailer
with tandem axle)

Truck Type-5
(16.8%)
Two-axle single unit truck (front and rear axle with dual wheel)

107
Vehicle Type
Illustration of its Axle (Short description)
(Occupancy)

Truck Type -6

(14.2%)
Three axle single unit truck (front and tandem)

Truck Type-11

(4.8%) Five axle truck with multiple trailer (tractor with single-axle, one semi-trailer with
single-axle, another trailer with two single-axles)

Truck Type-8

(3.5%) Three axle single trailer truck (tractor with single-axle, one-semi trailer with
tandem-axle)

Truck Type-14

(2.6%) Five axle multi trailer trucks (tractor with single-axle, one semi-trailer with single-
axle, another trailer with two single-axles)

Truck Type-4 (1.3%)

Three-axles single-unit trucks (front, single, single)

Truck Type-7

(0.4%)
Four-axles single-unit trucks (single and tridem)

108
The observed traffic data can further be categorized based on load spectra of each axle (i.e.,
steering, single, and tandem axles). Histograms and axle load spectra of steering, single and
tandem axle as observed from WIM measurements are plotted in Fig. 5.3. Due to the presence of
multi-peaks, these load spectra are fitted with mixed normal distributions, also termed as Gaussian
Mixture Model (GMM). For each histogram of axle load spectra, best fitted GMM is obtained; its
probability density function and distribution parameters are shown in Fig. 5.3. As these GMMs
are composed of more than one normal distribution, these can effectively capture the behavior of
empty and loaded trucks. Following these distributions, random samples are generated for each
axle type and added according to representative truck types (Table 5.1) for obtaining random
GVWs. For an example, truck type 5 composes of steering and single axles; hence, GVWs of truck
type 5 are generated by taking summations of randomly generated GVWs from GMMs
representing steering and single axles. The procedure is followed to generate random samples of
GVW for all different types of trucks, and finally 1024 (which equals to 210) samples of GVW are
composed accordingly to the proportions of various trucks as reported in Table 5.1. Generated
GVWs are used to calculate dynamic stress at critical locations of girders.

Figure 5.3 Statistics of GVW from different axle types; (a) Steering axle, (b) Single axle, and (c)
Tandem axle [Note: wi, µi, σi are parameters for ith axle representing, respectively, mixing
proportion, mean, and standard deviation of Gaussian Mixture Models, GMM]

109
5.2.3 Transient dynamic analysis considered vehicle-bridge coupled system

Literature stated that dynamic stress cycles in bridge girders under traffic loading are
strongly site-specific (Laman and Nowak 1996). To obtain the traffic-induced stress history of
girders, field measurements of displacements and strains using gauging instruments are necessary.
However, obtaining field measurement for each individual bridge is an expensive and time-
involved process. Thus, dynamic analysis of bridge systems under moving traffic has become a
reasonable alternative for assessing traffic-induced stress histories of bridge girders. The same is
adopted here.

The vehicle-bridge interaction (VBI) dynamics is composed by two sets of differential


equations of motion, one set for the moving truck and other set for the bridge deck. The two
subsystems, vehicle and bridge interact with each other through the force at contact points between
the bridge and axles. To synthesize the dynamic response of the bridge deck under moving load, a
semi-analytical method is used as an analysis technique in which closed-form equations for
flexural vibration of the bridge girder is solved for time-variant moving load. This method is
particularly useful for the VBI analysis in the current study since investigated bridges responded
within the elastic range under GVWs. Thus, this semi-analytical method helped to reduce the
computational time without compromising accuracy in the obtained solution. In the process, the
moving vehicle represents the fatigue load model. Using recorded WIM data, Chotickai and
Bowman (2006) demonstrated that the three-axle fatigue load model provided by the American
Association of State Highway and Transportation Officials or AASHTO (AASHTO 1990) may
notably overestimate the fatigue damage of short-to-medium span bridge girders. Alternatively, a
new three-axle fatigue truck model is proposed that can accumulate fatigue damage primarily from
two- to five-axle trucks on highways (Chotickai and Bowman 2006). This proposed fatigue-load
model is adopted here as it fits better for considered bridge girders and observed truck types.

The adopted fatigue load model (shown in Fig. 5.4) is composed of rigid bodies to represent
vehicle body and axles, and all components are connected through springs and dampers. The
analytical representation of this load model in the form of spring-mass-damper system is shown
below the load model in Fig. 5.4. The vehicle body has two degrees-of-freedom (DOFs) –
rotational (  ) and vertical (ys) displacements. Each of the three axles is represented with an
independent DOF, yti in the vertical direction in which i denotes the axle number. Suspension and

110
damping of the vehicle body are represented as Ksi and Csi, respectively. The same components for
axles are denoted as Kti and Cti. Table 5.2 summarizes the suspension and damping properties of
the load model as obtained from Paraskeva et al. (2017). This fatigue load model is assumed to
travel at a maximum speed of 80 km/h, which is the maximum speed allowed for trucks in the U.S.
highways. The bridge deck is idealized as a discretized simply supported Euler-Bernoulli beam
with mass per unit length of 𝑀𝐵 , damping coefficient 𝐶𝐵 , moment of inertia I, and Young’s
modulus E.

Figure 5.4 Three-axle fatigue load model and its analytical representation for the ensuing analysis

111
Table 5.2 Parameters associated with truck model for VBI analysis

Parameter Magnitude

L1 7.60 (m)

L2 2.10 (m)

L3 8.60 (m)

Rolling moment of inertia of truck body (J) 38200 (kg-m2)

Upper spring stiffness of first axle (Ks1) 485208 (N/m)

Lower spring stiffness of first axle (Kt1) 1,750,164 (N/m)

Upper spring stiffness of second axle (Ks2) 3806344 (N/m)

Lower spring stiffness of second axle (Kt2) 7006614 (N/m)

Upper spring stiffness of third axle (Ks3) 3938068 (N/m)

Lower spring stiffness of third axle (Kt3) 7014858 (N/m)

Upper damper coefficient of first axle (Cs1) 190000 (N-s/m)

Lower damper coefficient of first axle (Ct1) 98000 (N-s/m)

Upper damper coefficient of second axle (Cs2) 390000 (N-s/m)

Lower damper coefficient of second axle (Ct2) 190000 (N-s/m)

Upper damper coefficient of third axle (Cs3) 390000 (N-s/m)

Lower damper coefficient of third axle (Ct3) 190000 (N-s/m)

112
At any given location (x) and time (t, in sec), the governing differential equation of motion
of the moving fatigue load model and the flexural vibration of the bridge can be expressed as

[𝑀𝑡 ]{𝑦̈ 𝑡 } + [𝐾𝑡 ]{𝑦𝑡 } + [𝐶𝑡 ]{𝑦̇ 𝑡 } = {𝐹𝑡 } (5.1)

....
𝑁𝑎
𝑀𝐵 𝑦̈ 𝐵 (𝑥, 𝑡) + 𝐶𝐵 𝑦̇ 𝐵 (𝑥, 𝑡) + 𝐸𝐼 yB (𝑥, 𝑡) = − ∑𝑖=𝑖 𝑓𝑖 (𝑡)𝛿(𝑥𝑖 − 𝑣𝑡) (5.2)

where Mt, Kt, and Ct are, respectively, mass, stiffness, and damping matrices of the truck; 𝑦𝑡 , 𝑦̇ 𝑡 ,
and 𝑦̈ 𝑡 are, respectively, the displacement, velocity, and acceleration vectors of the truck; Ft is the
....
vector of wheel-bridge contact forces acting on the bridge; 𝑦̇ 𝐵 , 𝑦̈ 𝐵 , and yB respectively, the first,

second and fourth order derivatives of bridge deflection. Na is the number of axles in fatigue load
model, fi(t) is the time-variant contact wheel force from ith axle on the bridge deck,  ( xi − vt ) is

the Dirac delta function which is used to model the presence of ith axle at a distance x from the
bridge extreme end, and  is travelling velocity of the truck. The contact force fi(t) can be given as

𝑓𝑖 (𝑡) = 𝑊𝑖 − 𝐾𝑡𝑖 (𝑦𝑡𝑖 − 𝑦𝑐𝑖 ) − 𝐶𝑡𝑖 (𝑦̇ 𝑡𝑖 − 𝑦̇ 𝑐𝑖 ) (5.3)

where, Wi is the static gravity load of vehicle imposed on bridge deck at ith axle contact point, Kti
and Cti are stiffness and damping of ith axle, 𝑦𝑡𝑖 and 𝑦̇ 𝑡𝑖 are, respectively, vertical displacement and
velocity of ith axle, 𝑦𝑐𝑖 and 𝑦̇ 𝑐𝑖 are corresponding quantities at contact points.
In this analysis, the longitudinal unevenness or the vertical irregularity of road h(x) at each
location x along the bridge with span length L can be generated through a series of simple harmonic
functions as given in the equation next following Agostinacchio et al. (2014).

𝑁𝐻 𝑛
ℎ(𝑥) = ∑𝑗=0 √Δ𝑛2𝑘 10−3 (𝑗Δ𝑛
0
) cos (2𝜋𝑗 Δ𝑛𝑥 + 𝜑𝑗 ) (5.4)

According to the cited literature, Δ𝑛 expresses the discretized spatial frequency which can
be calculated as 1/L; k is a magnitude that depends on the classification of road according to ISO
8608 (2016); 𝑛0 is the spatial frequency which is adopted as 0.1 cycles/m; the maximum theoretical
sampling frequency nmax in sampling interval B can be calculated as 1/B, and NH is the number of
terms in the simple harmonic function which is used to build up the roughness and it is estimated
as nmax / Δ𝑛; and 𝜑𝑗 is the random phase angle which is assumed to follow uniform probabilistic

113
distribution within the range 0–2π. The classification for road surface profile adopted in the current
study is Class A as per ISO 8608 (2016), and necessary parameters according to this road class are
taken to model the surface elevation at every location on the bridge. In the past, comprehensive
studies are performed to explore the influence of road surface roughness on fatigue reliability and
service life of bridges. Exploring similar impacts, however, is beyond the scope of the current
study. As Class A road profile is considered in the study here and trucks are assumed to travel on
the bridge within their specified speed limit, any possible separation between vehicle wheels and
the road due to road bumpiness is ignored.

The time-dependent mid-span deflection of the bridge (𝑦𝐵 ) can be calculated using mode
superposition technique, which approximates the displacement response using linear combination
of vibration modes. Detailed formulation on this can be found in Saidi et al. (2017). Obtained
deflection is used to calculate stresses, 𝜎(𝑥, 𝑡) in rebars at a depth zg below the neutral axis of
girder cross-section. The analysis is repeated for 1024 stochastic samples of GVW ranging
between 55 kN to 310 kN (i.e., the lowest and highest observed GVWs). Fig. 5.5a shows stress-
time histories of fatigue-critical rebars for three different GVWs. As these plots indicate, the
maximum mid-span stress occurs when the steering axle is on the mid-span. Synthesized stress-
time history for each sample of GVW is statistically analysed to quantify an equivalent stress that
will be used further to estimate fatigue damage of reinforcing steel and concrete.

The discussed procedure for estimating deflection of girders under moving load is validated
using observed field measurements reported in Deng and Cai (2009). For this purpose, properties
associated with the vehicle and bridge are obtained from Deng and Cai (2009) to maintain
consistency. Fig. 5.5b shows a comparison between the measured response from that literature and
the same predicted through the discussed VBI model. Result shows a reasonable agreement
between the two, and hence, it confirms the suitability of the VBI model for the ensuing analysis.

5.2.4 Estimation of equivalent fatigue stresses

Obtained stress-time histories are composed of high cycle and low amplitude stresses (Fig.
5.5a). These stress histories contain many pulses, and thus, it becomes difficult to describe the
stress cycle directly that can be used for fatigue analysis. Hence, rainflow cycle counting method
is used first to extract constant stress ranges and associated frequencies from these complex stress-
time histories (Yan et al. 2017). Subsequently, S-N (stress-life) curves and the linear damage

114
Figure 5.5 Results from the VBI analysis; (a) mid-span stress-time histories of fatigue-critical
rebars of the bridge girder under three GVWs and (b) comparison of bridge deflection obtained
from in-situ field test and the current study.

accumulation hypothesis is applied to convert computed stress ranges for each passage of vehicle
(i.e., GVW) to a single equivalent stress. This equivalent stress is expected to produce fatigue
damage of rebars same as that would result from corresponding stress ranges (Yan et al. 2017).

In AASHTO (2008), the S-N curve for steel uses a single slope value of 3 to predict fatigue
life. However, available study demonstrated that the use of a single-slope S-N curve may result in
an underestimation of fatigue life due to the existence of variable low amplitude stress cycles
(Kwon et al. 2012) that are even lower than the constant amplitude fatigue limit or CAFL. The
CAFL comprises of stress range that corresponds to infinite fatigue life. To overcome this issue,
past research proposed the use of bilinear S-N curves which would be more appropriate than single-
slope curves for low amplitude stress cycles to estimate fatigue life (Yen et al. 2013). Accordingly,
the bilinear S-N curve from Eurocode 3 (CEN 2005) is used in the study here.

Since truck traffic loads are stochastic in nature, the stress ranges in stress spectrum have
variant amplitude of stresses and frequencies. It should be noted that corroded rebars contain initial
small degree of damage which is propagated further by fatigue stress amplitudes higher than
CAFL. As damage accumulates and propagates in rebars, they continue to get affected by stress

115
amplitudes even lower than CAFL (Kwon et al. 2012). With the use of Miner’s damage criterion
(Miner 1945) and S-N curve, the expression of equivalent stress ∆𝜎𝐸𝑞 is given as

nq  q3 nr  r5

 q  D KC
+ 
 r  D KD
 5
=
( n +  n )
Eq (5.5)
q r

KD

where ∆𝜎𝐷 represents CAFL, ∆𝜎𝑞 and ∆𝜎𝑟 are qth and rth stress ranges which are,
respectively, higher than or equal to ∆𝜎𝐷 and less than ∆𝜎𝐷 ; nq and nr are corresponding numbers
of stress cycles; KC and KD are coefficients to represent fatigue strength. Values of KC and KD
depend on the fatigue detail category as described in the Eurocode 3. For RC bridges, the fatigue
detail category of rebar is butted joint (Yan et al. 2017), and associated parameters are: ∆𝜎𝐷 = 59
MPa; KC = 1.02×1012 (MPa)3; KD = 1.64×1014 (MPa)5. This procedure is followed for every
generated sample of GVW, and 1024 values of ∆𝜎𝐸𝑞 are obtained. Fig. 5.6 plots histogram of these
equivalent fatigue stresses; it shows that these samples can be fitted into a mixed normal
distribution. The weighted average (or, mean) of ∆𝜎𝐸𝑞 is taken to evaluate damage imposed on
steel and concrete as vehicles traverse on the bridge. Though the mean ∆𝜎𝐸𝑞 is used in the ensuing
analysis, generation of 1024 samples was necessary to capture the contributions from all possible
truck types as per their observed relative occupancies on the freeway while estimating fatigue-
induced equivalent stresses. Gathered knowledge on the possible variation of ∆𝜎𝐸𝑞 will be useful
for a thorough uncertainty analysis considering key uncertain parameters pertinent to the real-life
corrosion exposure, structural parameters, and variable traffic load demand to gauge their
simultaneous impact on seismic vulnerability and resilience of aging RC bridges. The same will
be done as part of a future study.

For each girder, number of stress cycle per truck passage is set to 1 (Wang et al. 2016), and
the total number of daily truck passages is calculated using the parameter Annual Average Daily
Truck Traffic (AADTT). Thus, the total number of equivalent stress cycles in a year can be
represented by the volume (i.e., accumulation) of truck traffic, which is 365 times the observed
AADTT. The cumulative traffic cycles for the case study region over the observed period (i.e.,
between 1995 and 2001) is obtained from Lu et al. (2002). Based on the observed traffic growth

116
over the recorded period, linear and compound growth models (adopted from Lu et al. 2009) are
used to predict time-evolving trends for truck traffic in future years.

Figure 5.6 Histogram of equivalent fatigue stress of rebar under varying GVW

The linear growth model is termed as “conservative growth” and the compound growth
model is herein called as “aggressive growth”. Literature suggests that for highways in California,
4% yearly growth can be used to project truck traffic in future years (Lu et al. 2009). Accordingly,
cumulative projected truck traffic up until year 2094 (i.e., 100 years from 1995) for two growth
models are evaluated and plotted in Fig. 5.7. These cumulative traffic cycles are used to estimate
fatigue damage on the bridge girders.

5.2.5 Effect of corrosion-fatigue on properties of steel rebar

5.2.5.1 Corrosion of bridge girder and pier

To evaluate the impact of corrosion-fatigue on geometric properties of steel rebars, the


impact of corrosion is discussed first. Using the Fick's second law of diffusion, the study estimates
the time for corrosion initiation (Tini) for a structural member with cover depth equal to Xcov,
considering the exposure to deicing salt. This estimation is determined using the following
equation

117
Figure 5.7 Projected growth in truck traffic volume based on the observed traffic

2
𝑋𝑐𝑜𝑣 𝐶𝑠 −𝐶𝑐𝑟 −2
𝑇𝑖𝑛𝑖 = [𝑒𝑟𝑓 −1 ( )] (5.6)
4𝐷𝑐 𝐶𝑠

Here Ccr, Dc, and Cs, respectively, represent the critical chloride concentration for corrosion
initiation, diffusion coefficient, and chloride concentration at the surface of concrete. For the
bridges analyzed here, Xcov is taken as 50.8mm. A specific value of Ccr is difficult to obtain even
for a given region due to the spatial variability of environmental condition. Caltrans Corrosion
Guidelines (Caltrans 2021) express that Ccr may vary uniformly between lower and upper bounds
of 0.71 and 1.77 kg/m3, respectively. Likewise, a randomly generated value of 1.62 kg/m3 from
this Uniform distribution is used for Ccr in the current study. Note that as the distribution for Ccr is
Uniform, every generated sample within the given bound is equally probable. A mean value of
161.15 mm2/year is used for Dc which is obtained based on observations from field survey
containing 252 concrete samples from 49 bridges in CA (Weyers et al. 1994). Another field survey
on 16 concrete bridges in CA reported measured chloride concentrations at different depths of
bridge decks (Spellman and Stratfull 1969). Reported mean concentrations at 25.4mm and 50.8mm

118
depths are utilized, and the value of Cs is backtracked using Eq. (5.6). Estimated mean Cs is then
used to calculate Tini for of the current study, which is found to be 8 years for the investigated
bridges. The same values are taken for girders and piers of both bridges.

Assuming uniform corrosion of rebars, available gross sectional area of a rebar A(tz) after
z years from the time of corrosion initiation (Tini) shall be determined as given below in which dini
is initial (or, uncorroded) diameter of rebars and rcorr represents the rate of corrosion.
𝜋
𝐴(𝑡𝑧 ) = 4 [𝑑𝑖𝑛𝑖 − 𝑟𝑐𝑜𝑟𝑟 (𝑡𝑧 − 𝑇𝑖𝑛𝑖 )]2 𝑓𝑜𝑟 𝑡𝑧 > 𝑇𝑖𝑛𝑖 (5.7)

Here an average value of 0.05 mm/year is adopted for rcorr as suggested in the Caltrans
Corrosion Guidelines (Caltrans 2021). Note that any secondary effect of corrosion (such as
reduction in yield strength of steel and compressive strength of concrete, and loss of bond) are not
considered here to manage the computational challenges involved in projecting the impact of
material degradation due to corrosion-fatigue to the life-cycle seismic resilience of RC bridges.
Further, past research found that uniform mass loss of rebars due to uniform corrosion does not
result in any notable change in stress-strain response of rebars in tension (Kashani et al. 2013).
Thus, the possibility of inelastic buckling of corroded rebars is ignored here by taking uniform
stiffness distribution along the length of corroded rebars owing to the uniform corrosion of rebars.

5.2.5.2 Corrosion-fatigue of bridge girder

Fatigue behavior of rebars is mainly governed by the stress range and maximum stress limit
(Kim & Heffernan 2008). After the onset of corrosion, effective stress range of rebars starts
increasing in an accelerated rate under fatigue load cycles (as corrosion results in reduced cross-
sectional area of rebars). This phenomenon leads to the development of fatigue cracks on rebars,
which result in further reduction of rebar cross-sectional area with time (Zhang et al. 2017). Thus,
the time-dependent reduction of the cross-sectional area of a steel rebar during fatigue life would
indicate the time when fatigue damage may occur in corroded steel bars. Following Zhang et al.
(2017), the residual area Asf ( N ) at which the corroded rebars get fracture can be expressed as

As
Asf ( N ) =  s ,max (5.8)
fy

119
where N is the number of cycles which caused the corroded rebar to get fracture at residual area
Asf ( N ) , and it is also called as fatigue life of rebar;  s ,max is the maximum nominal stress on rebar;

f y is the yield strength of corroded rebar, and As is the cross-sectional area of rebar at the time

when corrosion initiates. Thus, traffic-induced high cycle and low amplitude stresses can lead to
steel fracture of corroded rebars. This is also called elastic fatigue of steel rebars as induced stresses
are lower than yield strength of steel. Past studies demonstrated that with the accumulation of
stresses over time in steel rebars due to number of stress cycles from truck traffic, rebars may
experience a linear loss in cross-sectional area with number of stress cycles till it reaches to the
fatigue life Nf (Song et al., 2019). The reduction of rebar area can be formulated based on Fig. 5.8,
at any time ti since initiation of corrosion, the reduction in rebar area which underwent n1 number
of traffic cycles can be calculated as a sum of fatigue and corrosion phenomenon. It is expressed
as

A1 = As − 1c − 1f (5.9)

where, 1c is change in rebar area due to corrosion; and 1f is the change in rebar area due to fatigue.

Similarly, the reduction in area of rebar which underwent n2 number of traffic load cycles at t2
years is given as

A2 = As −  c2 − ( 1f +  2f ) (5.10)

Therefore, the time-dependent reduction in area of rebar due to the combined impact of
corrosion-fatigue is expressed generally as

  m n −n 
Ai ( t , n ) = ( d − (t − Tini ) rcorr ) −  i i −1 ( Ai −1 − AfN ) 
2
(5.11)
 i =1 Ni −1 − ni −1
i
4 
where, Ai ( t , n ) represents the residual area of rebar at time t and exposed to n number of traffic

load cycles. Based on the experimental study, Zhang et al. (2012) proposed an analytical
expression of fatigue life N of naturally corroded rebar having  s degree of corrosion as given in

Eq. (5.12).

log N = (14.282 + 17.029 s ) − ( 3.319 + 8.743 s ) log  Eq (5.12)

120
Figure 5.8 Schematic representation of corrosion-fatigue deterioration of embedded rebar in
girder

Fatigue failure will occur if the number of applied traffic cycles nz exceeds fatigue life Nz.
To demonstrate the effect of traffic growth on fatigue damage of rebar, the mean equivalent stress
obtained from Eq. (5.8) is substituted in Eq. (5.14) to get the time varying fatigue life of rebars.
Palmgren-Miner’s rule suggests that each stress range with nz cycles of repetition contributes to
the fatigue failure by nz/Nz. The fatigue failure for such repetition of stress cycles will occur if the
ratio nz/Nz reaches to one. The fatigue damage to rebar for two varying traffic growth models is
compared in the left quadrant of Fig. 5.9; it shows that the aggressive growth model accumulates
damage on rebar at a high rate as compared to the other model and takes the ratio to 0.7 at 100th
year age of the bridges. It is important to note in this relation that due to low amplitude of obtained
fatigue stresses and stress ranges (which are in accordance with that reported in Yan et al., 2017),
the impact of corrosion degradation on these stresses is not studied in the current research.

For the two traffic growth models, impact of corrosion-fatigue on cross-sectional area of
steel rebar is estimated using Eq. (5.11-5.12) and presented in Fig. 5.9a. Note that 1024 random

121
samples of GVW resulted in variable area reductions (shown with bands), from which the current
study considers the mean estimates as portrayed with solid lines in Fig. 5.9a

Figure 5.9 Time-dependent reduction of (a) residual area of rebar due to corrosion-fatigue and
(b) elastic modulus of concrete due to fatigue [Uncertainty bands (mean±1 SD) represent the
variability of mean estimates]

Though fatigue-induced damage in concrete is less than that for steel, the combined
degradation in steel and concrete may become significant for the performance of bridge girders
under traffic load. The total concrete strain 𝜀𝑐𝑡 under fatigue load is the sum of the elastic strain
𝜀𝑐𝑒 and strain increment 𝜀𝑐𝑝 (plastic strain) due to repetitive load. Following Han et al. (2015), 𝜀𝑐𝑝
is given as
1 1
3
𝜀𝑐𝑝 = 129𝜎𝑐,𝑟𝑚 𝑡𝑡𝑟𝑢𝑐𝑘 + 17.8𝜎𝑐,𝑟𝑚 ∆𝜎𝑐,𝑟 𝑛𝑧3 (5.13)
(𝜎𝑐,𝑚𝑎𝑥 ±𝜎𝑐,𝑚𝑖𝑛 )
Here 𝜎𝑐,𝑟𝑚 is the relative mean stress of concrete that can be estimated as
2𝑓𝑐𝑘

where 𝜎𝑐,𝑚𝑎𝑥 and 𝜎𝑐,𝑚𝑖𝑛 are, respectively, the maximum and minimum stresses in concrete; nz is
the number of traffic loading cycle at time tz; ttruck is the time duration of truck passage in hours;
(𝜎𝑐,𝑚𝑎𝑥 −𝜎𝑐,𝑚𝑖𝑛 )
∆𝜎𝑐,𝑟 is the relative stress range of concrete which can be estimated as . The variation
2𝑓𝑐𝑘

of 𝜎𝑐,𝑚𝑎𝑥 with respect to total strain can be approximately represented using the apparent elastic
𝑓
modulus of concrete (𝐸𝐶 ) under nz loading cycle as given below (Balaguru 1981).

122
𝑓 𝜎
𝐸𝑐 = 𝜀 𝑐,𝑚𝑎𝑥 (5.14)
𝑐𝑒+𝜀 𝑐𝑝

Fig. 5.9b presents the ratio of change in modulus of concrete to its original modulus for
two different traffic growth models. It can be observed that the aggressive traffic growth model
results in slightly higher deterioration of concrete properties than that from the other growth model.

5.2.5.3 Effect of corrosion-fatigue on stiffness of bridge deck

Reductions in cross-sectional area of rebars and elastic modulus of concrete result in


reductions in effective moment of inertia of bridge girders and flexural stiffness of bridge decks
with time. These are plotted in Fig. 5.10 for both traffic growth models. Stiffness-based approach
is followed to evaluate degraded stiffness of bridge decks under corrosion-fatigue. As can be seen
from Fig. 5.10a, percent reductions in moment of inertia for both traffic growth models follow the
trends observed in Fig. 5.9a for the area reduction of rebars along bridge age. The percent reduction
in deck stiffness shown in Fig. 5.10b, however, is a combination of reductions in moment of inertia
and elastic modulus of concrete; thus, Fig. 5.10b shows higher reduction values with time when
compared to that for moment of inertia. For the same reason, the influence of aggressive traffic
growth model over the other growth model gets prominent in Fig. 5.10b much prior to that in Fig.
5.10a.

Figure 5.10 Time-dependent reduction of (a) moment of inertia of a T-girder and (b) flexural
stiffness of deck

123
5.3 Application to representative highway bridges

5.3.1 Representative bridges and finite element models

The influence of material deterioration owing to corrosion-fatigue on seismic performance


of highway bridges is demonstrated on two representative multi-span simply supported T-girder
RC bridges having two and three 20-meter-long spans. Both bridges are considered to have the
same traffic, environmental, and hazard exposure conditions. Hence, obtained results for these two
bridges help to understand how the impact of corrosion-fatigue on bridge seismic performance
may change based on bridge flexibility. Each span is composed of five parallel T-girders with
intermediate diaphragms. These girders are supported on column bents consisting of three 914.4-
mm-diameter circular columns. Adjacent spans at column bent are supported on elastomeric
bearings. Bridge columns has 12 numbers of #29 longitudinal rebars. Both ends of the bridge girder
has seat-type abutments with elastomeric bearings. Bridge columns are supported by pile
foundations, each of which contains eight piles. Fig. 5.11 represents schematics of the three-span
bridge and its three-dimensional (3D) finite element (FE) model at the pristine condition. The two-
span bridge is a replica of this three-span bridge except for the middle span and one column bent.

FE models of these bridges are developed in OpenSees. To account for the damage owing
to corrosion-fatigue and its impact on constitutive behaviour of steel and concrete, a fiber-section
model is assigned for RC bridge girders. The model was developed by utilizing 3D nonlinear
beam-column elements with the necessary number of fibers and integration points that relate the
element force with constitutive models of composing materials. Deck slabs are represented as
grillage of elastic elements and required sectional properties are obtained from discretized cross
sections. Intermediate diaphragm within the bridge span is represented as rigid elastic elements in
the transverse direction. To ensure rigid action between bridge girders and decks, girder and deck
elements are constrained together by rigid elements. Therefore, developed 3D models of bridge
girders are assemblages of nonlinear and linear elements that are constrained together with rigid
elements at respective joints. This method of developing 3D FE model of bridge girders was
adopted from available literatures that developed FE model of existing concrete and steel bridges
for the nonlinear time history analysis (Tapia-Hernández et al. 2017; Wan et al. 2021). The
schematic FE model of the bridge girder in Fig. 5.11 shows the major linear and nonlinear
elements.

124
Figure 5.11 Three-dimensional finite element model of the three-span bridge with linear and
nonlinear elements

Suitability of the superstructure model is validated by using observational data from a full-
scale test conducted by Song et al. (2002) on an in-situ T-girder bridge. The same superstructure
is modeled in OpenSees following the described procedure and analyzed under the load used for
the field experiment. Fig. 5.12 shows a comparison between simulated load-deflection response of
the girder at the mid-span with the same observed from the experiment, and a reasonable agreement
is observed between the two. Hence, the developed grillage model can effectively capture the
distributed stiffness of the bridge superstructure.

Other bridge components (such as piers, abutments, bearings, and pile foundations) are
modelled following the approach prescribed in Yilmaz et al. (2016). Schematics of these elements
are presented in Fig. 5.11 and a short discussion is presented here. Readers may refer to Chapter 3
(Section 3.3) for a detailed discussion on each of these nonlinear elements. Bridge piers are
modelled using displacement-based nonlinear fiber elements, and properties of associated
materials are described in fibers. The nonlinear behavior of steel and concrete is simulated using
Concrete07 and Steel02 material models in OpenSees. Such a modeling approach for bridge piers

125
is validated by Yilmaz et al. (2016) and followed in several past studies focusing on the seismic
performance evaluation of bridges. The structural behavior of the bridge foundation is idealized
by using translational and rotational springs in the respective directions. To represent expansion
gaps in the longitudinal direction of abutments, elastic-perfectly plastic gap elements are used. In
the transverse direction, abutment stiffness is modeled through elastic-perfectly plastic elements
(without gap). In two horizontal translational directions, bearings on abutments and column bent
are modelled using elastic-perfectly plastic elements. Pounding between adjoining decks and deck-
abutment interfaces are captured through contact elements as described in Muthukumar (2003).

Time-dependent reduction of rebar area (uniform along rebar length) and elastic modulus
of concrete of RC girders due to corrosion-fatigue for each life-cycle year (as obtained from Fig.
5.9) are implemented to the FE models of bridges. Moreover, area reduction of longitudinal rebars
of piers due to corrosion is accounted for while updating FE models of bridges for every life-cycle
year. Thus, these FE models capture deterioration of investigated bridges along their service lives.

Figure 5.12 Comparison of load-deflection response of the bridge girder obtained from in-situ
test and the current study

126
5.3.2 Nonlinear time history analyses of aging bridges

Nonlinear time-history analysis of investigated bridges, at pristine and degraded states, are
performed for 30 sets (i.e., 60 motions) of orthogonal ground motions (GMs) having 2%, 10% and
50% exceedance probabilities in 50 years (Somerville et al. 1997). These ground motions,
originally generated for the California region, are used here as information related to bridge
corrosion and traffic load is also acquired for the same state. For each set of orthogonal GMs,
response time-histories of critical bridge components (discussed next) are recorded. To observe
the influence of traffic growth models on bridge seismic response, maximum longitudinal
displacements of bearings at pier bents of the three-span bridge at its 100-year service life are
shown in Fig. 5.13 for some of the earthquakes recorded at different stations. As this figure clearly
demonstrates the severity of the aggressive traffic growth over the other, seismic fragility and
resilience of the investigated bridges under aggressive traffic growth are shown henceforth for the
demonstration purpose.

Figure 5.13 Seismic response of bearings for different traffic growth models

127
5.3.3 Impact of corrosion-fatigue on bridge seismic vulnerability and seismic resilience of
bridge

For seismic vulnerability analysis of the investigated bridges, recorded seismic response
quantities of different bridge components are expressed in terms of EDP, following a method
similar to the one discussed in Chapter 3. The process is followed for different degraded conditions
of the bridges till they reach 100 years of lifespan. Fig. 5.14 shows seismic fragility curves of the
three-span bridge at four damage states and for five life-cycle years. Corresponding median values
of fragility curves are mentioned within the figure. As this figure shows, almost no variation in
fragility curves is observed at the minor damage state of the bridge. Besides, bridge fragility curves
degrade indicating increasing seismic vulnerability of the bridge as deterioration proceeds. Similar
trends are observed in seismic fragility curves developed for the two-span bridge. These time-
dependent fragility curves are used further to estimate life-cycle seismic resilience of these bridges.

5.3.4 Impact of corrosion-fatigue on bridge seismic resilience

Seismic resilience of bridges provides an integrated measure of bridge seismic damage,


losses, and recovery. The resilience value indicates how quickly the damaged infrastructure can
get back to its pre-event functionality level. Theoretical background for estimating seismic
resilience of bridges is discussed in detail in previous publications by the authors (Banerjee et al.
2019; Vishwanath and Banerjee 2019; Devendiran et al. 2021). The same framework is used here.
Post-event seismic losses, direct and indirect, are estimated based on the fragility information
derived in the preceding section of this article. Estimated losses are integrated with appropriate
damage-state specific recovery functions (as developed in Vishwanath and Banerjee 2019) to
estimate resilience of the investigated bridges. The resilience value is computed as a dimensionless
quantity between 0 and 100%. Higher value indicates better resilience.

Fig. 5.14 shows seismic resilience of the two-span and three-span bridge, respectively, at
pristine and various degraded conditions for the aggressive traffic growth model. The assessment
of resilience for different intensity of earthquake is performed using the framework presented in
Chapter 3. It can be observed that with increasing intensity of ground motions (i.e., PGA),
resilience of the bridge reduces. For every PGA value, seismic resilience of the bridges reduces as
deterioration proceeds. Obtained results suggest that even though both bridges have almost the
same seismic resilience at the pristine condition, the three-span bridge experiences higher

128
reduction in resilience due to aging. To measure the rate of resilience reduction with the level of
bridge deterioration, percent reductions in resilience with respect to resilience values at the pristine
state of respective bridges are calculated.

Maximum reductions of 10.4% and 13.1% in seismic resilience are observed for the two-
span bridge, respectively, at the 75th year and 100th service life. These corresponding quantities for
the three-span bridge are 11.1% and 14.1%. Certainly, these values can have wide uncertainty
bands if uncertainties from all possible sources are considered in the analysis framework.
Nevertheless, observed results imply significant adverse impact of the combined corrosion-fatigue
degradation on life-cycle seismic performance and resilience of RC bridges.

Figure 5.14 Time-dependent fragility curves of the three-span bridge at (a) minor damage, (b)
moderate damage, (c) major damage, and (d) collapse state (values within parenthesis represent
median values

129
Figure 5.15 Life-cycle seismic resilience of the aging bridges; (a) two-span bridge and (b) three-
span bridge

5.4 Closure

In summary, the study highlights the adverse impact of corrosion-fatigue degradation,


resulting from the combined interaction of chloride-induced corrosion and traffic-induced stress
cycles, on the life-cycle seismic resilience of RC bridges. The study estimated the impact of this
degradation by performing seismic analysis on two RC T-girder bridges with different spans. The
obtained results reveal declining trends in parameters such as the cross-sectional area of embedded
steel rebars in girder, elastic modulus of concrete, moment of inertia of girders, and flexural
stiffness of bridge decks. These trends are a consequence of the accumulated damage arising from
corrosion-fatigue degradation. The observed deterioration in constituent materials over time
increases the seismic vulnerability of the bridges, leading to reduction in overall resilience
throughout the bridge service life. Specific observations from the study provide quantitative
insights into the effects of corrosion-fatigue degradation on various aspects of bridge performance.
The comparative assessment between the two bridges suggests that the impact of corrosion-fatigue
on seismic resilience becomes more complex as bridges become more flexible due to a higher
number of spans. The same effect is expected for bridges with longer span lengths, even if the
number of spans remains unchanged. Overall, the results emphasize the significant role of
corrosion-fatigue degradation in determining the long-term performance of RC bridges. Ignoring

130
the impact of corrosion-fatigue degradation reduces the seismic safety and resilience of bridges
throughout their lifespan. Implementing appropriate strategies and considering uncertainties in the
analysis framework are crucial for mitigating the confronting effects of corrosion-fatigue
degradation and maintaining the seismic safety of bridges over their lifespan.

131
This page is intentionally left blank

132
CHAPTER 6

ANALYTICAL ASSESSMENT OF VERTICAL GROUND MOTION


EFFECT ON HIGHWAY BRIDGE UNDER THE IMAPCT OF
CORROSION AND TRAFFIC INDUCED FATIGUE

In previous chapter, the importance of accounting deterioration in bridge deck due to


corrosion-fatigue is emphasized. This form of deterioration is modelled as uniform dissolution of
metal for assessing the seismic vulnerability and resilience of aging bridges located in urban areas.
This chapter presents an improved model that considers the coupling phenomenon of pitting
corrosion of rebars and fatigue cracks in rebars to model the residual area of rebars due to
deterioration owing to corrosion-fatigue. Such degradation results in reduction in structural
stiffness of bridge deck, making it vulnerable to earthquake ground motions, particularly for the
ones which have the active participation of VGM. The influence of VGM has not been explored
yet for the seismic vulnerability assessment of highway bridges. The main objective of this chapter
is to discuss and demonstrate the confronting role of corrosion-fatigue degradation and VGM on
seismic response of various bridge components through a representative multi-span simply
supported RC bridge in Gujarat, India. Seismic fragility models for bridge components such as
girders, elastomeric bearings, piers, and abutments affected by corrosion-fatigue under the
combined influence of vertical and horizontal ground motions are presented as fragility curves.

6.1 Case-study region and a representative bridge

Gujarat, situated in the western region of India, is one of the most earthquake-prone areas
despite being away from plate boundaries. The occurrence of 2001 Bhuj earthquake (Mw 7.7) in
the Kachchh region of Gujarat and the presence of several active faults in the region indicate the
potential of this region to experience devastating earthquakes in future. It has been reported that
the 2001 Bhuj earthquake has resulted in 13819 casualties and an economic burden exceeding
USD 10 billion (Rastogi et al. 2001). In comparison to other states in India, Gujarat region is
unique as it encompasses all recognized seismic zones (V, IV, III, II), based on the seismic zoning

133
map of India (BIS 2002). The region is a tri-junction of failed rift (large separation of earth crust);
Kachchh, Cambay and Narmada with several active-faults (Biswas 1987). Some the identified
active faults in the region are Nagar Parkar fault (NPF), Allah Bund fault (ABF), Island Belt fault
(IBF), Kachchh mainland fault (KMF), South Wagad fault (SWF) and Katrol Hill fault (KHF).
The tectonic features of Gujarat, illustrated in Chopra et al. (2012), is adopted for the study here
and presented in Fig. 6.1.

Figure 6.1 Tectonic features of Gujarat (Adopted from Chopra et al. [2012])

The seismic risk in Gujarat is likely to increase due to rapid expansion of its major urban
centers and industrial zones. The region facilitates refineries, chemical industries, marine
infrastructures such as ports and harbor. The rapid expansion of industries leads to a rise in the
demand of more truck traffic on highway bridge of Gujarat, which may lead to fatigue condition
to bridge decks. The state of Gujarat is also influenced by environmental factors that affect the
corrosion potential of metals. Some of the identified factors that influence the corrosion potential
in Gujarat are humidity, exposure to salt spray and other damaging species, and variation in
temperature. These two scenarios can potentially result in corrosion-fatigue condition to embedded
rebar in RC girders and affects the overall structural performance under seismic events.

To demonstrate the impact of corrosion-fatigue condition on seismic performance of


highway bridges, a multi-span simply supported T-girder representative bridge which is assumed

134
to be located in Gujarat, India, is taken for this purpose. Multi-span T-girder bridges are very
common in Gujarat, and they comprise of nearly 80% bridges in this region (Dey and Sil 2021).
Fig. 6.2 depicts basic dimensions of the bridge and cross-sectional details of its components. The
bridge design meets the specifications of Indian Roads Congress (IRC:112-2011). The bridge
consists of three 20-meter-long spans with 11-meter-wide deck. Each span is composed of five-
parallel T-girders which are interconnected through intermediate diaphragms. The bridge has two
pier bents, each having three 914.4mm diameter piers. Each column has 12 numbers of #32
longitudinal rebar (area of longitudinal reinforcement = 805 mm2) and #16 circular stirrups spaced
at 150 mm center-to-center along the column length. On two sides, the bridge deck is supported
on seat-type abutments with elastomeric bearings. Additional elastomeric bearings are placed on
top of pier. Bents are supported on pile foundations.

Figure 6.2 Baseline dimension of representative T-girder bridge

135
6.2 Intensity and vertical seismic component of 2001 Bhuj earthquake (Mw 7.7)

The epicenter of Mw 7.7 Bhuj earthquake 2001 was Bhacau (a town which is 80 km away
from Bhuj in Gujarat). This devastating earthquake which lasted up to 110 seconds, destroyed
buildings, killed people, and collapsed major road infrastructures leading to Bhuj. The recorded
acceleration time-history of the seismic event in three translation direction is illustrated in Fig. 6.3.

Figure 6.3 Acceleration time history in (a) N-S direction (b) E-W direction (c) U-D direction

Ground motions which are generated close to the source of earthquake event are typically
characterized by large amplitude and high frequency vibration. For near-field earthquakes, the
ratio between the peak ground acceleration (PGA) of vertical component to horizontal component
can be greater than one (Kim et al. 2011). The vertical-to-horizontal spectral ratio for the 2001
Bhuj earthquake is represented in Fig. 6.4a which shows that the ratio varies depending on the
spectral period. At the lower period range, the ratio exceeds the code specified limit, and even
crosses 1.0 mark indicating that the code-specified two-third rule can result in an unconservative
seismic design of bridges with low fundamental periods. The figure also compares the response
spectra, Fourier amplitude spectra of vertical and horizontal ground motion of 2001 Bhuj
earthquake. In terms of the energy content, VGM in the near-field domain has majority of energy
content at its short period range. This can be observed from the response spectra of seismic event
(shown in Fig. 6.4b). The spectral acceleration of both vertical and horizontal components almost
equals in the short period range. The vertical component of seismic event can cause uplift or
settling of bridge, whereas the horizontal component can impose lateral forces such as shear and
bending. Hence, at short period range, these two components can couple and cause significant
damage to structure. Fig. 6.4c illustrates the comparison of fast Fourier transform (FFT) of

136
horizontal and vertical component of the seismic event. The plot shows the dominant frequency of
VGM lies between 5 Hz – 10 Hz indicating that the VGM has accumulated the majority of its
energy in a narrow high frequency band. This can cause damage to bridges whose vertical vibration
mode falls within this range of frequencies.

Figure 6.4 (a) Vertical-to-horizontal spectral ratio of 2001 Bhuj earthquake which recorded at
Ahmedabad station (b) Spectral acceleration of horizontal and vertical component of seismic
event (c) Fourier amplitude spectrum horizontal and vertical component of 2001 Bhuj earthquake

According to Ambraseys and Simpson (1996), Ambraseys and Douglas (2003) and
Ramadan et al. (2021), the ratio of spectral accelerations of VGM to HGM is influenced by site to
source distance along with the local soil condition. Due to the local site effect, different regions of
Gujarat experience different intensity of earthquakes. The field damage evidence from the 2001
Bhuj earthquake by Narayan et al. (2002) reported that structures located 150 km away from the
epicentre of the quake experienced axial failure of columns and wall upliftment as a result of strong
vertical shaking. The V/H ratios of other observed aftershocks of this earthquake in Gujarat with
different epicentral distance and soil condition are listed in Table 6.1. The V/H ratio of an
accelerogram recorded in Bachau station (which has an underlying rock deposit) at an epicentral
distance of 39 km is 0.73, whereas the same recorded at another station located at an epicentral
distance of 65 km is 1.01. Similarly, the V/H ratio of motions recorded at stations on soil deposit
are observed to be higher for greater epicentral distance. This observation clearly illustrates the
importance of considering VGM to quantify the impact of strong vertical shaking even though the
structure is located away from faults. This has a specific significant for aging structures.

137
Table 6.1 Observed peak ground acceleration (PGA) of aftershocks of 2001 Bhuj earthquake
(data obtained from Choudhury et al. 2016)

Epicentral Recorded PGA (g)


Station name distance Geology type V/H
(km) NS EW UD

Bachau 39 Rock 0.12 0.10 0.08 0.73

Anjar 65 Rock 0.08 0.06 0.07 1.01

Dayapar 141 Rock 0.14 0.10 0.06 0.51

Kandla 69 Soil 0.22 0.09 0.06 0.43

Mundra 115 Soil 0.06 0.02 0.06 1.73

Radhanpur 137 Soil 0.04 0.02 0.05 1.77

6.3 Estimation of corrosion-fatigue damage on steel and concrete under site-specific


traffic load

6.3.1 Observed truck traffic and its axle load statistics

The information of truck traffic load on the representative bridge is obtained from the
recorded WIM measurements from a highway corridor of Gujarat (MES 2022). This WIM data
contains information from an axle load survey of heavy vehicles that are monitored for a period of
24 hours. The data set includes axle loads and gross vehicle weights (GVWs) of 335 units of trucks
(both empty and fully loaded).

The obtained GVW is plotted as scatter plot and presented in Fig.6.5, with the x-coordinate
represent the numeric order of the observed truck, and y-coordinate representing the magnitude of
GVW. Description of these trucks and their percentage occupancies in the observed sample are
listed in Table 6.2. As can be seen from the table that the multi-axle truck with dual wheel on either
side (three axle truck) dominates the composition with 41% occupancy, followed by two axle truck
with tandem wheel on either side with 33%, and two axle truck with dual wheel on either side with
23%, and two axle truck with single wheel on either side with 3% occupancy.

138
Figure 6.5 Observed GVW of truck in a highway corridor of Gujarat

Table 6.2 Truck type and its percentage occupancy in the monitored samples

Truck type I II III IV

Single axle
Single axle with Tandem axle with Multi axle with dual
with single
Description dual wheel on dual wheel on wheel on either side
wheel on
either side either side
either side

Percentage
3% 23 % 33 % 41 %
occupancy

Illustration
(top view)

139
Each axle load spectra in the WIM dataset are classified as steering axle, single axle, or
tandem axle. Fig. 6.6 depicts the histogram of observed axle load spectra from WIM measurements
for each axle type. The presence of empty and fully loaded trucks in the sample has resulted in
multiple peaks in the histogram, which are fitted using the mixed normal distribution, also known
as the gaussian mixture model (GMM). The best fit for each histogram in Fig. 6.5 is obtained, and
its statistical properties are inscribed within the figure. Based on the observed data as presented in
Table 6.2 and Fig. 6.6, random samples of GVW for the four representative trucks are generated
by drawing random samples from each GMM and adding them according to the percentage
occupancy of trucks. For an example, Truck type IV is composed of steering, single, and tandem
axles; thus, random axle weight from GMM representing steering, single, and tandem axles are
drawn and added together to get the GVW of the representing truck. Once many such samples of
four truck types are generated, a final pool of 1024 truck (i.e., GVW) are obtained according to
their individual occupancy percentages. In the pool, the highest and lowest GVWs are,
respectively, 435 kN and 50 kN, as per the observed samples. Note that the generation of more
truck samples helped in obtaining a thorough distribution of GVW between its upper and lower
observed bounds, which further facilitates in obtaining well distributed values of traffic-induced
stresses. More discussion is presented later.

Figure 6.6 Statistics of observed axles in the sample – (a) Steering axle (b) Single axle (c)
Tandem axle

140
6.3.2 Corrosion-fatigue of rebar

The damage mechanism of rebar subjected to corrosion-fatigue is conceptually represented


in Fig. 6.7. The entire mechanism gets started from the ingress and diffusion of chloride ions
through concrete cover. Corrosion of embedded steel rebars initiates when chloride concentration
at the concrete-steel interface reaches to a threshold limit. The process is then followed by the
formation of corrosion pits on the surface of rebar due to non-uniform corrosion. Because of this
non-uniform nature, the depth of pit may vary along the length as well as the cross-section of
rebars. In case when the pit formation is nearly uniform across rebar cross-section (i.e., uniform
loss of cross-sectional area), it can be referred to as uniform corrosion; otherwise, it is called pitting
corrosion (Section A-A and B-B in Fig. 6.7). Naturally, rebars become weak at locations of pitting
corrosion and may develop fatigue cracks when stresses accumulate owing in rebars to repetitive
traffic load cycles. These cracks grow with time and may reach to a critical crack size causing
rebars to fail in brittle fracture and resulting in reduced strength of bridge girders. Prior reaching
to the critical crack size, formation of fatigue cracks also reduces the efficient cross-sectional area
of rebars for resisting traffic loads on the girders, which in turn gives rise to higher stresses in
rebars. Thus, the reduction in cross sectional area of rebars should account for the impact of
equivalent stresses from traffic and frequency of repetitive truck traffic loads, in addition to the
effect of corrosion. The following section discusses the consequences of corrosion and corrosion-
fatigue phenomena in detail. In previous studies, the reduction of rebar area due to deterioration
due to corrosion-fatigue is deduced as a uniform form of metal loss (Zhang et al. 2017; Song et al.
2019). However, it has been experimentally observed that fatigue cracks initiate from the location
of imperfection (or corrosion pits) and induce brittle fracture once they reach a critical size. As a
result, it is necessary to numerically model the degeneration of steel area while accounting for the
pitting mechanism.

When the exposed surface of girder comes into the direct contact with chloride laden
environment, chloride molecules diffuse through the capillary pores of concrete cover, initiate
corrosion and depassivate the steel. The diffusion of chloride ions can be idealized mathematically
using the Fick's second law of diffusion. This idealization applies only to concrete surfaces that
are completely immersed in and permanently saturated with chloride species. However, complete
saturation of RC members is not always the case, and so, the applicability of the diffusion equation
for every practical scenario becomes doubtful. To overcome this, an error-function is introduced
141
Figure 6.7 Conceptual representation of corrosion-fatigue damage mechanism of rebar

within the formulation of 1D diffusion to approximate the estimated initiation time for any field
exposure condition (Bertolini et al. 2004). DuraCrete (2000) also introduced several correction
coefficients to realistically estimate the efficient diffusion rate of chloride species in concrete
continuum for atmospheric exposure scenario. The widely utilized analytical model for evaluating
the initial time of corrosion 𝑇𝑖𝑛𝑖 in concrete members located in chloride laden environments is
given as
1
𝐶2 𝐶 −𝐶 −2 (1−𝑛𝑐𝑙 )
𝑇𝑖𝑛𝑖 = {4𝑘 𝑘 𝐷 (𝑡 )𝑛𝑐𝑙 [𝑒𝑟𝑓 −1 ( 𝑠 𝐶 𝑡ℎ )] } (6.1)
𝑒 𝑐 𝑐𝑙,0 0 𝑠

where, 𝐶 is the cover depth (in mm) of the exposed specimen, 𝐶𝑠 is the surface chloride
concentration on the specimen, 𝐶𝑡ℎ is the threshold concentration of chloride at which rebar
corrosion initiates, and 𝑒𝑟𝑓 −1 is the inverse Gaussian error function. 𝑘𝑐 and 𝑘𝑒 are correction
coefficients, respectively, for curing and environmental factor. 𝐷𝑐𝑙,0 is the reference diffusion
coefficient (mm2/year) that can be determined from a compliance test, which involves measuring
the change in volume of a concrete sample at t0 days as a function of concentration of diffusing
species and 𝑛𝑐𝑙 is the age exponent of which is used to express the susceptibility of reinforcement
to corrosion due to cement densification by hydration. Validity of Eq. 6.1 is confirmed by using it
to predict observed chloride concentrations at various depths of concrete panels of 1.0×0.5×0.12m
that were placed by Costa and Appleton (1995) in an outdoor chlorinated environment for 3 to 5
years. Analytical equation gave successful prediction.

The environment around Gujarat is hypersaline with shallow and salty marshes during the
monsoon, dry and salty flat land during the winter, and crystalline and salty atmospheric condition
during the summer (Pandit et al. 2015). The case study bridge is subjected to the atmospheric

142
exposure condition, and for the case study region, atmospheric chlorine concentration is observed
to have spatial distribution between 8000 ppm and 11000 ppm (Dey and Sil 2021). Based on this
reported range, an average value of 9500 ppm is adopted here as Cs. There are several attributes
that affect the magnitude of 𝐶𝑡ℎ for corrosion in RC members; these include concentration of
hydroxide ion in pore solution, electrochemical potential of steel, water-cement ratio, availability
of oxygen on the rebar surface, and pH of concrete. Because of that, no single value of 𝐶𝑡ℎ could
be decided from those reported in the published literature. The study here, therefore, takes the
value of 𝐶𝑡ℎ from Caltrans standard provisions of corrosion guidelines (Caltrans, 2021). Mean
values of kc, ke, 𝐷𝑐𝑙,0 , and 𝑛𝑐𝑙 are taken from DuraCrete (2000), and the values are 0.16, 0.68, 473
mm2/year, and 0.65, respectively. Applying the values of these parameters in Eq.6.1, corrosion
initiation time (Tini) for components (with 50.8mm or 2-inch cover depth) of the case study bridge
is estimated to be 7.7 years.

After the onset of corrosion, time-dependent rate of corrosion rcorr at tp years (since
initiation) is obtained from Vu and Stewart (2000) for the atmospheric exposure condition and
utilized here for life-cycle analysis. Note that rcorr is expected to decrease along the service life of
the bridge. Following Val and Melchers (1997), the pit depth p(t) at a time t in the bridge service
life can be estimated as a function of rcorr as
𝑡
𝑝(𝑡) = 𝑅 ∫𝑇 𝑟𝑐𝑜𝑟𝑟 (𝑡𝑝 )𝑑𝑡𝑝 (6.2)
𝑖𝑛𝑖

where R denotes the pitting factor, which is the ratio of maximum pit depth to average depth caused
by uniform corrosion. During the oxidizing period, this factor will approach to one, indicating that
the uniform form of corrosion has progressed deeper. The value of R in this study is adopted from
Stewart and Al-Harthy (2008). The time to nucleation of a pit is estimated by solving the integral
function by defining p(t) in Eq.6.2 with a threshold pit depth of 1.98 µm (Bastidas-Arteaga et al.
2009). Based on the adopted parameters, the time to pit nucleation for the case study bridge is
roughly estimated to be 8 years. When the formation of pit reaches to a threshold depth, it marks
the locations for fatigue cracks to initiate as a result of repetitive traffic loading.

143
6.3.3 Traffic-induced fatigue stresses

The procedure and the essential mathematical model which required to synthesis the
dynamic stresses of the representative T-girder bridge under moving truck load are explained in
Chapter 5 (refer Section 5.2.3). Following the same concept explained in Section 5.2.4, stress-time
history from each sample of truck in converted to a single equivalent stress which is expected to
produce the same level fatigue damage to bridge girders as that under the given stress time history.
The procedure is followed for all 1024 samples of GVW and obtained 1024 equivalent stresses are
plotted with corresponding GVWs in Fig. 6.8a. As observed, equivalent stress increases almost
linearly with GVW. These equivalent stresses are used to obtain the combined impact of corrosion
and fatigue on performance degradation of girders. As shown in Fig. 6.8b, the obtained equivalent
stresses are fitted with the appropriate number of Gaussian distributions. The fitted distribution is
a mixture of three Gaussian distributions with mean values of 2.39 MPa, 4.49 MPa, and 7.45
MPa and with a mixture proportion of 0.25, 0.28, and 0.47, respectively.

Figure 6.8 (a) variation of equivalent stress with GVW (b) Gaussian mixture model of equivalent
stress

144
6.3.4 Deterioration of rebar due to corrosion-fatigue accounting pitting and fatigue cracks

In the existing literature, pit depth on rebar surface is modelled as cup-shaped


hemispherical pits (Stewart 2004; Ghosh and Sood, 2016; Shekhar et al. 2018). Following that, the
degradation of rebar due to combined corrosion and fatigue phenomenon is modelled in this study
as a hemispherical area loss. The deterioration mechanism of pristine rebar under corrosion-fatigue
is conceptually illustrated in Fig. 6.9. In this, the three plots refer to (a) uniform corrosion of rebar,
(b) condition of deep pit (DP) with fatigue cracks following pit nucleation, and (c) corrosion-
fatigue (CF) degradation. Fatigue damage in rebars increases with the growth and propagation of
fatigue cracks, and results in a reduced efficient cross-sectional area of rebars. Following Bastidas-
Arteaga et al. (2009), the rate of crack growth under a given stress intensity factor and frequency
of applied truck load is estimated as
𝑑𝑎
= 𝐶𝑝 (𝛥𝐾)𝑚 𝑓 (6.3)
𝑑𝑡

where, a is the crack length, 𝛥𝐾 is the alternating stress intensity factor at corrosion pits,
𝐶𝑝 and m are fatigue material constants of embedded steel rebars, and f is frequency of cyclic load
(expressed here as number of trucks per day). Values of 𝐶𝑝 and m are dependent on the
aggressiveness of the environment (Bastidas-Arteaga et al. 2009). Eq.6.3 is an extended
formulation of the Paris-Erdogan law. In this equation, the stress intensity factor 𝛥𝐾 at corrosion
pits of rebar with initial diameter D is quantified in terms of pit depth 𝑝(𝑡) and stress range 𝛥𝜎 as
given below.
𝑝(𝑡)
𝛥𝐾 = 𝛥𝜎𝑌 ( ) √𝜋𝑝(𝑡) (6.4)
𝐷

Here,  is the equivalent fatigue stress (obtained from Fig. 6.8a) and Y(𝑝(𝑡), 𝐷) is a
dimensionless factor for the notched geometry of the specimen and can be expressed
(approximately) as given below (Bastidas-Arteaga et al. 2009).

𝑝(𝑡) 𝑝(𝑡) 2 𝑝(𝑡) 3 𝑝(𝑡) 4


𝑝(𝑡) 1.121−3.08( )+7.344( ) −10.244( ) +5.85( )
𝐷 𝐷 𝐷 𝐷
𝑌( )= 3 (6.5)
𝐷
𝑝(𝑡) 2 2
[1−2( ) ]
𝐷

Calculated pit depth (using Eq. 6.2) at each year is considered to be an initial imperfection
in rebars, and the change in geometric notched factor accounting for the formation of pit in
estimated using Eq. 6.5. Estimated value is used in Eq. 6.4 to calculate stress intensity factor 𝛥𝐾
145
at a tip of corrosion pit for a given equivalent stress . Here, three mean equivalent stresses from
the GMM in Fig. 6.8b are taken as . Finally, fatigue crack length for the associated stress
intensity factor for each time period t is calculated using Eq. 6.3. In this calculation, load frequency
(f) of traffic load is taken as 335 trucks/day following the WIM measurement data on annual daily
traffic (ADT) of trucks for the case study region (MES 2022). The value of f is expected to increase
with time following the normal traffic growth. The data on heavy truck production volume in India
from 2011 to 2021 (Statisca 2023) indicates that the average compound growth of transport
vehicles (including light and multi-axled trucks) has increased 5% over the last decade. Following
the statistics, a traffic growth rate of 5% is considered here to extrapolate the observed load
frequency of traffic in future years. This information is utilized to estimate time-dependent residual
area of rebars under corrosion-fatigue.

Figure 6.9 Demonstration (schematic) of area loss of rebar due to (a) uniform corrosion alone,
(b) the formation of deep pit (DP) and fatigue cracks after pit nucleation at pitting locations, and
(c) effective residual area of rebar under combined corrosion and fatigue

For pitting corrosion, various methods including electrochemical techniques and image
analysis are followed in practice to evaluate the residual area of rebar. As the current study deals
with combined corrosion-fatigue phenomenon, the geometry-based approach suggested by Val
and Melchers (1997) is used here to quantify the residual area of rebar. Here, 𝑤(𝑡) represents the
width of the pit after pit nucleation and the formation of fatigue cracks (shown in Fig. 6.9b).
Corresponding residual cross-sectional area of the rebar is 𝐴𝐷𝑃 (𝑡). The pit depth 𝑝𝐶𝐹 (𝑡) due to the
combined corrosion-fatigue deterioration can be estimated as sum of depth of deep pit 𝑝(𝑡) and
crack length 𝑎(𝑡) [i.e., 𝑝𝐶𝐹 (𝑡) = 𝑝(𝑡) + 𝑎(𝑡)]. Thus, the growth of pit width 𝑤(𝑡) due to the
growth of corrosion pit is estimated as given below.

146
𝑝𝐶𝐹 (𝑡) 2
𝑤(𝑡) = 2𝑝𝐶𝐹 (𝑡)√1 − ( ) (6.6)
𝐷

Subsequently, the time-dependent residual area of rebar ADP is quantified as (following Val and
Melchers 1997)
D
𝐴𝑖 − (𝐴1 − 𝐴2 ) for 𝑝𝐶𝐹 (𝑡) ≤
√2
𝐴𝐷𝑃 (𝑡) = { (6.7)
D
(𝐴1 − 𝐴2 ) for D > 𝑝𝐶𝐹 (𝑡) ≥
√2

where, 𝐴𝑖 represents the area of pristine rebar (shown in Fig. 6.9a) and

𝑝𝐶𝐹 (𝑡)2 𝑤(𝑡)


𝐴1 = 0.5 [𝜃1 (0.5𝐷)2 − 𝑤(𝑡) |0.5𝐷 − |] and 𝜃1 = 2𝑎𝑟𝑐𝑠𝑖𝑛 [ ] (6.8)
𝐷 𝐷

2 𝑝𝐶𝐹 (𝑡)2 𝑤(𝑡)


𝐴2 = 0.5 [𝜃2 (𝑝𝐶𝐹 (𝑡)) − 𝑤(𝑡) ] and 𝜃2 = 2𝑎𝑟𝑐𝑠𝑖𝑛 [2𝑝 ] (6.9)
𝐷 𝐶𝐹 (𝑡)

As previously demonstrated, corrosion-fatigue degradation of rebar accounts for both


uniform and pitting corrosion along rebar length (Fig. 6.7c and Fig. 6.7d). Thus, the effective cross-
sectional area of rebar due to corrosion-fatigue 𝐴𝐶𝐹 (in Fig. 6.9c) is derived as (upon modifying
the expression given by Ghosh and Sood (2016) to estimate residual area of rebar under pitting
corrosion alone)
𝑤(𝑡)
𝐴𝐶𝐹 (𝑡) = [1 − ] [𝐴𝑈 (𝑡) − 𝐴𝑖 ] + 𝐴𝐷𝑃 (𝑡) (6.10)
2𝐷

where, 𝐴𝑖 and 𝐴𝑈 (𝑡) are, respectively, the pristine area of rebar before corrosion and residual area
of rebar at a time t after corrosion initiates under the assumption of uniform corrosion (Fig. 6.9a).
Obtained ratios of 𝐴𝐶𝐹 and 𝐴𝑖 for rebars in the given bridge girder under three equivalent stresses
are shown in Fig. 6.10a. To demonstrate the effect of fatigue on rebar deterioration, area ratio for
only corrosion scenario (without corrosion-fatigue mechanism) is also plotted there. As can be
seen from the figure, the residual area of rebar drops up to 47%, 16% and 8% with respect to the
pristine condition in case corrosion-fatigue degradation under 7.5, 4.5, and 2.4 MPa equivalent
fatigue stresses, respectively. The same is only 6% under the corrosion alone condition. Though
this corrosion alone degradation is almost at the same level of corrosion-fatigue degradation under
2.4 MPa fatigue stress, differences are significant for equivalent stresses greater than 2.4 MPa.
These higher equivalent stresses are expected to be from truck type III and IV that have higher
occupancy percentages within the recorded WIM data (as per Table 6.2).

147
Fatigue causes degradation in concrete properties as well, although it is not as significant as
that in the steel rebar (refer to Section 5.2.5.2). Nevertheless, corrosion-fatigue deteriorations in
both concrete and steel properties are accounted for to estimate the reduction in flexural stiffness
of girders over bridge service life (plotted in Fig. 6.10b). It can be observed that the flexural
stiffness of bridge girder drops up to 14.8%, 10.8% and 6.7% from the pristine condition in case
of corrosion-fatigue degradation under equivalent fatigue stresses of 7.5, 4.5 and 2.4 MPa,
respectively. The same for corrosion alone is 5.9%.

Figure 6.10 Reductions in (a) residual area of steel rebar and (b) flexural stiffness of bridge girder

6.4 Finite element model to account the effect of VGM on bridge structure

6.4.1 Elastomeric bearing

T-girders are supported on elastomeric bearings that transfer superstructure load to


bents and abutments. During a seismic event, elastomeric bearings resist the shear demand through
sliding and develop complex nonlinear response. To capture this sliding behavior, elastomeric
bearings are provided with some initial shear stiffness such that these elements can resist shear
force until the coefficient of friction (µ) at the interface of elastomer and concrete exceeds its
limiting value. After reaching the limit (i.e., when shear force = µ × NBearing; NBearing is the normal
force), the stiffness of the elastomeric bearing drastically reduces and becomes close to zero. This
behaviour is idealized using the Steel01 material in OpenSees.

148
In previous studies on seismic performance evaluation of bridges considering VGMs,
bearing stiffness was assumed to be constant (Aryan and Ghassemieh 2020). However,
experimental observations revealed that the coefficient of friction for elastomeric rubber bearings
can be expressed as a function of variable normal stresses on bearings resulted from VGM of the
seismic event (Steelman et al. 2013). In a more recent study by Xiang and Li (2018), coefficient
of interfacial friction µ during a seismic event is estimated in terms of the normal stress, σ (in MPa)
on the bearings and the sliding velocity, v (in mm/sec) as given below.

µ = 1.09𝜎 −0.59 − (1.09𝜎 −0.59 − 1.02𝜎 −0.72 )𝑒 −0.39𝑣 (6.11)

Xiang and Li (2018) validated the expression using experimental data. Further validation
is done for the current study with normal stresses σ = 2.0 and 4.0 MPa and sliding velocity v = 0.5
mm/s. As can be understood, the shear strength of the elastomeric bearing decreases when normal
stress exerted on the bearing relaxes. It indicates the correlation between gravity load and shear
response of elastomeric bearings, and the same is used in this article to demonstrate the influence
of VGM on the material behavior of elastomeric bearing. In two principal horizontal directions,
material behavior of elastomeric bearings is assigned with nonlinear zeroLength spring elements
in OpenSees (Fig. 6.11). During seismic analysis with VGM, time-histories of normal force and
sliding velocity exerted on elastomer are recorded and corresponding 𝜇𝑒𝑐 for each time instance
of the ground motion is calculated. The parameter and updateParameter commands in OpenSees
are used to calculate and update the shear strength of an elastomer (i.e., µ × NBearing) at each time
instance of the seismic event having VGM. In the vertical direction, elastomeric bearings are stiff
in compression but not rigid. This vertical stiffness of bearings is idealized using the elastic-no-
tension (ENT) material model, and its value is calculated following Warn et al. (2016).

6.4.2 Pier

Flexural behavior of bridge piers under atmospheric exposure scenario for uniform and
pitting corrosion is modelled in OpenSees using dispBeamColumn element. The modelling
strategy used by Ghosh and Sood (2016) for degraded bridge piers due to pitting corrosion is
adopted here. In addition to the flexure behavior, bond-slip behavior of bridge piers during seismic
events is also accounted for in the numerical modeling. The anchorage slip of longitudinal rebar
during a seismic event may cause rigid-body rotation of piers causing additional deformation to
piers. In OpenSees, zeroLength rotational springs with linear constitutive relationship are utilized

149
to approximate this behavior. The linear relation expresses the stiffness of the rotational spring
(Krot) as a function of the instantaneous axial load. Mathematically, the relation is given as follows
(Elwood and Eberhard 2006).
8𝑢𝐸𝐼𝑒𝑓𝑓
𝐾𝑟𝑜𝑡 = (6.12)
𝑓𝑦 𝐷𝑙

Here, u represents average tension between the surface of concrete and longitudinal rebar
and can is approximated as 0.5√𝑓𝑐𝑘 , where fck is the characteristic compressive strength of
concrete. fy and Dl are, respectively, the yield strength and diameter of longitudinal rebar. EIeff
expresses the effective flexural rigidity of piers that can be calculated as the function of
instantaneous axial load P which is exerted on piers during the seismic event.
𝑃
0.2 𝐸𝐼 𝑖𝑓 ≤ 0.2
𝐴𝑔 𝑓𝑐𝑘
5𝑃 4 𝑃
𝐸𝐼𝑒𝑓𝑓 = 𝐸𝐼 (3𝐴 − 3) 𝑖𝑓 0.2 ≤ 𝐴 ≤ 0.5 (6.13)
𝑔 𝑓𝑐𝑘 𝑔 𝑓𝑐𝑘
𝑃
0.7 𝐸𝐼 𝑖𝑓 0.5 ≤ 𝐴
{ 𝑔 𝑓𝑐𝑘

In this equation, E, I and Ag are the elastic modulus, moment of inertial, and gross cross-
sectional area of a pier, respectively. In the numerical simulation in OpenSees, time-history of the
instantaneous axial force (P) exerted on piers under seismic excitations is recorded, and
corresponding Krot is calculated. This calculation and subsequent update of Krot in the numerical
model of the bridge are performed at each time instance of a seismic event using parameter and
updateParameter commands in OpenSees.

6.4.3 Bridge deck and other components

The bridge deck is represented as a grillage of various element types which are available
in OpenSees. The proposed modelling approach in this chapter employs the same method as
discussed in Chapter 5 (refer Section 5.3.1). The FEM of the representative bridge is illustrated in
Fig. 6.11. The material behaviour of a seat-type abutment is modelled in the FEM using zeroLength
non-linear translational springs. In the longitudinal direction, the passive stiffness of soil offered
to the back-wall of abutment, as well as the passive and active stiffness of pile offered to the
abutment are modelled using Hysteretic materials. The material behaviour of soil and pile stiffness
to bridge can be combined using a parallel material object. The stiffness contributed by the pile is

150
the transverse direction is modelled using Hysteretic material. The stiffness of the abutment and
pile in necessary directions are assigned in accordance with the recommendations of Caltrans
(SDC 2019). Foundations are modelled and idealized using simple lumped linear elastic springs
that can support both translational and rotational behaviour in all degrees-of-freedom (DOF) of
foundations.

6.5 Selection of ground motion for dynamic analysis (based on frequency content of 2001
Bhuj earthquake)

Selection of ground motion is an important step in assessing the dynamic response of civil
structures. While assessing the vulnerability and risk of bridges in a region, it is desirable to use
input ground motions that realistically represent the regional seismic potential. For seismic
analysis of the bridge in Gujarat, number of recorded accelerograms for this region is extremely
low that imposed a challenge to conduct the current study. To overcome this, additional ground
motions in accordance with the seismicity of Gujarat region is obtained from the PEER database
(2023). During reconnaissance of the 2001 Bhuj earthquake, reverse fault was found to be focal
mechanism (Antolik and Dreger, 2003). For all recorded past earthquakes (including Bhuj) in this
region, the range of moment magnitude Mw is found to lie between 5.3 and 7.7 (Mandal et al.
2003). Thus, these two selection criteria related to the fault mechanism and Mw are used for initial
screening of ground motions from the PEER database. Further down selection of initially screened
ground motions is done based on the frequency content of the 2001 Bhuj earthquake as detailed
next.

Dynamic response of a structure gets amplified when the frequency content of earthquake
ground motion matches with the fundamental vibration frequency of the structure. This is why
determining the frequency content of an earthquake ground motion is important while selecting it
to generate input ground motion data suite. A variety of indicators are currently being used to
characterize the frequency-content of earthquake ground motions. Among these, mean time period
(Tm) of ground motion is widely accepted measure to estimate the frequency content of earthquake
ground motions (Ratje et al. 1988). Following Ratje et al. (1988), Tm is expressed as
1
∑𝑖 𝑐𝑖2 ×( )
𝑓𝑖
𝑇𝑚 = ∑𝑖 𝑐𝑖2
𝑓𝑜𝑟 0.25 𝐻𝑧 ≤ 𝑓𝑖 ≤ 20 𝐻𝑧 (6.14)

151
Figure 6.11 Representative bridge and its nonlinear elements

where, Ci and fi represent, respectively, the coefficient of Fourier amplitude and discrete
fast Fourier transform (FFT) frequency of accelerograms spaced within the domain of 0.25 Hz to
20 Hz. The accelerograms of the 2001 Bhuj earthquake are first converted to frequency domain,
and then FFT analysis is performed to obtain ci at each fi, and then Tm is estimated following Eq.
6.14 for all three components of this earthquake. Thus obtained three Tm values for three
translational components of the Bhuj earthquake are used as final reference periods for ground
motion selection. For all down-selected ground motions from the PEER database (after the initial
screening), Tm values for the three translational components are estimated and compared with
respective component obtained for the Bhuj earthquake. Here, 15% tolerance is maintained for
each individual component. Following the process, a total of six ground motions are selected and
listed in Table 6.3. Finally, ground motion amplitudes are influenced by the site soil condition.
Thus, the final set of ground motions (in Table 6.3) are scaled up, without altering V/H ratios, with
suitable multiplying factors to match up the design spectra of the Gujarat region. The time interval
between the arrival of PGA of HGM and VGM is also studied (Table 6.3). Positive time lag
indicates that PGA of HGM occurs before VGM in the seismic event, and vice-a-versa. A time lag
of 0.01sec or 0.12 sec indicates that PGAs for both HGM and VGM occur concurrently.

152
Table 6.3 Ground motions used for the nonlinear time history analyses of the case study bridge

Focal Tm Tm Tm Time
GM Earthquake name (Mw,
mechan (sec) (sec) (sec) V/H lag
ID year)
ism (N-S) (E-W) (U-D) (sec)

#1 Bhuj (7.7, 2001) Reverse 0.598 0.668 0.347 0.66 2.88

#2 San Fernando (6.6, 1911) Reverse 0.702 0.781 0.352 0.69 0.01

#3 Friuli (6.0, 1976) Reverse 0.587 0.704 0.390 0.70 0.12

#4 Tabas (7.3, 1978) Reverse 0.581 0.703 0.386 0.75 1.30

#5 Tabas (7.3, 1978) Reverse 0.512 0.580 0.400 0.45 0.22

#6 Coalinga (6.4, 1983) Reverse 0.583 0.697 0.379 0.32 4.71

6.6 Impact of corrosion-fatigue and VGM on life-cycle response of bridge

6.6.1 Impact on natural frequency of bridge

Modal parameters such as natural frequency and mode shapes are subjected to change
owing to structural damage and degradation. To demonstrate the change in dynamic behaviour of
the bridge due to corrosion-fatigue degradation, girder and pier elements are updated with
appropriate mass and cracked stiffness values as the bridge ages, and modal analyses of the bridge
at different life cycle year are performed. Fig. 6.12 illustrates fundamental mode shapes (i.e., 1st
longitudinal, 1st transverse, 1st twisting, and 1st vertical) of the bridge at its pristine and 100-year
degraded conditions. Corresponding time periods are also mentioned in the figure. As expected,
diameter reduction of rebars due to corrosion in piers and corrosion-fatigue in girders made the
bridge flexible from its pristine to the 100-year degraded condition. Between these two health
conditions of the bridge, its fundamental time periods in the longitudinal, transverse, twisting, and
vertical modes of vibration have reduced by 0.3%, 1.2%, 1.2%, and 35%, respectively. These
reductions, particularly in the vertical mode of vibration, attribute to amplify bridge seismic
response when VGM component, in addition to HGM components, of an earthquake is considered
for seismic analysis. Results presented further sections elaborate on this.

153
Figure 6.12 Fundamental mode shapes and time periods of the bridge at pristine and 100-year
conditions

6.6.2 Impact on seismic response of bridge girders

The change in seismic response of the girder due to degradation is first demonstrated
through the resonant frequency of girder acceleration. With the decrease in girder stiffness due to
corrosion-fatigue (as shown in Fig. 6.10b), resonant frequency decreases. The fast Fourier
transform (FFT) of acceleration of bridge girder in the vertical direction under GM # 1 is shown
in Fig. 6.13a, in which peaks represent the resonant frequency of girder acceleration. As observed,
the resonant frequency reduces from 5.24 Hz at the pristine state to 3.96 Hz at 100th year degraded
state of the bridge. Further, reduced flexural stiffness due to corrosion-fatigue results in increased
mid-span vertical displacement of girders. Fig. 6.13b illustrates the variation of mid-span girder
displacement (vertical) of the bridge at its pristine state and 100-year age under the same ground
motion. As observed, initial displacement of the girder (primarily due to gravity load) increases
from 30mm to 78mm as a result of bridge degradation. Moreover, deterioration of embedded rebar
causes significant loss in flexural capacity of girders. The moment capacity of the investigated RC
154
T-girder is evaluated following the Indian Standard code of practice for plain and reinforced
concrete (IS-456, 2000). The estimated value is 5296 kN-m for the pristine state of the girder. The
same capacity becomes 4512, 3103, and 1700 kN-m, respectively, for fatigue stress of 2.48, 4.48,
and 7.50 MPa at the end of 100 year service life when corrosion-fatigue degradation is accounted
for. In case of corrosion alone scenario, the capacity after 100 year will be 4746 kNm.

Figure 6.13 Response of a bridge girder under GM # 1 at the pristine and 100-year degraded
states; (a) fast Fourier transformation of girder acceleration in the vertical direction, (b) mid-span
vertical displacement, and (c) moment demand and estimated capacity

Besides corrosion-fatigue degradation, the impact of VGM on moment demand at the mid-
span of the girder under the same earthquake is shown in Fig. 6.13c. This figure also plots the
moment capacity of the girder at its pristine and 100-year degraded states. As the plot depicts,
consideration of VGM gives rise to the maximum mid-span bending moment by 41% with respect
to that when only HGM components are considered. This moment demand considering VGM
largely exceeds the capacity of the girder at its 100-year degraded state indicating a complete
failure of the degraded girder under this earthquake. Similar observations are obtained for other
ground motions as well. Table 6.4 lists moment demands for all six ground motions. It shows that
seismic events with V/H ≥ 0.66 can cause deck failure due to flexure (as moment demand >
capacity) at 100-year bridge service life, except for GM # 3. The same damage is not observed for
seismic events with V/H less than 0.66. Moreover, for V/H ≥ 0.66, moment demand ratios are
getting higher than 1.0 indicating the notable impact of vertical ground motion components.

155
6.6.3 Impact on seismic response of elastomeric bearing

Change in girder displacement due to corrosion-fatigue degradation affects the seismic


performance of elastomeric bearings that support girders at both ends. With the increase in vertical
displacement of girders, elastomeric bearings may get displaced from their initial positions, and
this initial displacement of a bearing determines its overall seismic performance. Fig. 6.14a and
Fig. 6.14b show polar plots of recorded maximum translational displacements of elastomeric
bearings placed on the top of an exterior and interior pier, respectively, in a pier bent. For both
bearings, maximum seismic displacement increases with bridge degradation.

Table 6.4 Comparison of moment capacity and demand at the mid-span of bridge girder

Moment capacity Moment demand (kNm) Moment

GM ID V/H at 100-year service demand ratio


HGM and Only HGM
life (kNm) 𝐕𝐆𝐌+𝐇𝐆𝐌
VGM ( )
𝐇𝐆𝐌

#1 0.66 1700 2168.1 1536.8 1.41

#2 0.69 1700 1725.6 1523.1 1.13

#3 0.70 1700 1655.5 1518.6 1.09

#4 0.75 1700 1714.1 1518.8 1.12

#5 0.45 1700 1580.1 1522.1 1.03

#6 0.32 1700 1409.5 1524.8 0.92

For the six ground motions considered, 12% and 11% higher displacements are recorded,
respectively, for the exterior and interior bearings due to bridge degradation over 100 years.
Among these two bearings, the observed difference in recorded displacements under any specific
ground motion also depends on the time lag between PGAs of HGM and VGM components of that
seismic event. Ground motions with a time lag more than one second (i.e., for GM #1, 4, and 6;
Table 6.3) produced relatively higher translational displacement of exterior bearings than the
interior ones, while other ground motions recorded either equal or less displacements for the
exterior bearing than the interior one.

156
Figure 6.14 Maximum translation displacements of elastomeric bearings at the pristine and 100-
year degraded states of the bridge; (a) bearing on exterior pier in a bent (b) bearing on interior
pier in a bent

The observed difference between the two lines in each plot of Fig. 6.14 is the sole impact
of bridge degradation. To portray the combined impact of bridge degradation and VGM, Fig. 6.15a
and Fig. 6.15b illustrate time-histories of normal stresses on bearings, placed on a pier bent above
the exterior and interior piers, under GM #1 with and without including the VGM component. As
can be seen, the normal stress on bearing varies within 1.92 - 9.25 MPa and 3.18 - 8.78 MPa,
respectively, by including and excluding VGM component for the exterior bearing (Fig. 6.15a).
The same ranges are, respectively, 2.55 - 6.22 MPa and 4.30 - 4.80 MPa in Fig. 6.15b for the
interior bearing. The higher range of variation of axial stress for the exterior bearing is due to
overturning moment of bridge girder about its longitudinal axis. This is the reason why exterior
piers experience higher axial forces compared to the interior piers.

Further, variation in compressive stress on bearings influences the coefficient of friction at


the interface of elastomer and bridge girder. Following Eq. 6.11, friction coefficient changes based
on axial stress on the bearing and the sliding velocity during the seismic shaking. Thus, higher
fluctuation in compressive stress on bearings can produce higher variations in the frictional
coefficient and resulting frictional force at the interface of girder and bearing. This phenomenon
is demonstrated in Fig. 6.15c by capturing the variation of friction coefficient of bearing (placed

157
on the interior pier) under the same ground motion. This figure shows that the friction coefficient
fluctuates both sides within a range of 0.32 to 0.63 approximately between 30 to 50 second when
VGM is included in the analysis, whereas the same has only limited variation in case when VGM
is excluded. This result is in accordance with the observed variation in axial force on this bearing
as plotted in Fig. 6.15b

Figure 6.15 Variations in bearing response under GM # 1; (a) compressive stresses on an exterior
bearing, (b) compressive stresses on an interior bearing, and (c) coefficient of friction of an
interior bearing

Variations in friction coefficient and normal force on bearing alter its shear-deformation
behavior. The variable shear strength model considered in this study (discussed in Section 6.4.1)
can capture the change in normal stress on bearings and related change in yield strength of
bearing caused by instantaneous girder acceleration in both upward and downward directions.
6.16a shows the hysteretic shear-deformation relation of an elastomeric bearing (above the interior
pier) obtained from the dynamic analysis of the pristine bridge with and without including VGM.
It can be observed that the shear demand of the bearing under the combined influence of HGMs
and VGM has increased by 14.1% than that when VGM is not considered. This enhanced shear
demand will transfer to piers resulting in a potential increase in shear deformation of piers. 6.16b
illustrates the shear demand of the underlying pier (interior) for the same earthquake; it can be
seen that the maximum shear demand attained by the pier for the dynamic event without VGM is
419 kN, whereas the same demand for the same dynamic event with VGM is observed to be 486
kN (i.e., 16% increase as compared 419 kN). This increase in shear demand in the pier is solely
transmitted by the elastomeric bearing due to the presence of VGM component in the seismic
shaking. Thus, ignoring the impact of normal stress variation on the bearing owing to VGM

158
components will result in inaccurate estimation of forces exerted on bridge components and their
maximum response values. This will underestimate the seismic demand from ground motions with
significant vertical components, and thus, overestimate seismic fragility characteristics of bridges.

Figure 6.16 (a) Shear-deformation relationship of elastomeric bearing above the interior pier and
(b) variation of shear demand imposed on the underlying pier under GM # 1

6.6.4 Impact on seismic response of bridge pier

In addition to the higher variation in shear demand, VGM imposes a continuous variation
of axial force on bridge piers (Kim 2008; Wang et al. 2013). Such variation in axial force can alter
the yield curvature of piers. Moreover, higher flexibility of bridges due to gradual degradation
yields lower yield curvature of piers. Analyzing the case study bridge, some slight reductions in
the yield curvature of the pier at various axial load levels are observed from the undegraded to the
degraded state of the bridge as plotted in Fig. 6.17a. Here, the axial load ratio refers to the
proportion of axial load exerted on the pier during a seismic event to the axial load exerted due to
gravity of superstructure. It shows (almost) linearity between axial load ratio and pier yield
curvature (Kim 2008). Further, gradual degradation and VGM can jointly impact the curvature
ductility demand of piers. Fig. 6.17b and Fig. 6.17c show two sets of hysteretic plots on moment-
curvature ductility demand (M-CDD) obtained under GM # 1 with and without inclusion of VGM
in the dynamic analysis of the bridge at the pristine and 100-year degraded states. Gradual
degradation of the bridge results in lower yield curvature but higher curvature demand on piers;
hence, curvature ductility demand (CDD) of the pier increases significantly from its pristine to the
degraded state. As illustrated in Fig. 6.17b, the maximum curvature ductility demand of the pier

159
at the pristine state is observed to be 5.27, which becomes 8.81 at the 100-year degraded state.
These values are obtained when only HGM components of GM # 1 are applied. The same values
become 6.20 and 9.77, respectively, in the presence of VGM component of GM # 1 (Fig. 6.17c).
Thus, these enhanced CDDs from 5.27 to 6.20 at the pristine state and from 8.81 to 9.77 at the
degraded state are due to the higher seismic demand from the ground motion when VGM
component is considered. It further implies higher seismic vulnerability of bridge piers in the
presence of VGM.

For all six ground motions considered in this study, CDDs for the four cases are shown in
Fig. 6.18 in a comprehensive manner. The maximum increase of 53% in CDD from analyses with
and without VGM is observed for GM # 6. This increase in CDD is case specific and may be
attributed to the time lag observed for various earthquakes. However, no definite conclusion can
be drawn here based on the observations from a limited number of seismic cases. Nevertheless,
curvature ductility of piers is observed to increase at 100-year service life when compared to the
pristine state. This increase in demand due to gradual degradation of the bridge is widely observed
in the literature.

Figure 6.17 Pier response at the pristine and 100-year degraded states; (a) yield curvature under
variable axial load ratios, (b) curvature ductility demands (CDDs) without the inclusion of VGM
component of GM # 1, and (c) curvature ductility demands (CDDs) with the inclusion of VGM
component of GM #1

As demonstrated in Fig. 6.16a, additional shear demand arises in piers due to the variation
in axial load resulted from VGM component. The same variation of axial force can as well change
the shear capacity of pier section. In this study, shear strength of a pier section is calculated using
the equation proposed by Priestley et al. (1994) In contrast to other available shear strength models
160
(such as ACI 318-05 (2005), AASHTO (2017), and Eurocode 2 (2004)), the one proposed by
Priestley et al. (1994) accounts for contributions of constituent materials (i.e., concrete and steel
reinforcement) and the axial compression of piers in resisting shear. Estimated shear capacity using
the model, therefore, gets influenced by the consideration of VGM and gradual degradation due to
corrosion-fatigue in two possible ways. Firstly, piers are expected to have higher fluctuations in
axial force (both tension and compression) when VGM is considered; this directly affects the shear
capacity of piers. Secondly, shear-resisting capacity of concrete depends on curvature ductility of
pier section; with VGM and higher degradation of piers, curvature ductility of piers increases
(shown in Fig. 6.17b and Fig. 6.17c) resulting in a decrease in the contribution of concrete towards
shear capacity of pier sections.

Furthermore, shear capacity of a pier section may have some impact from the difference in
arrival time of PGA between HGMs and VGM (i.e., time lag). As previously reported, increase in
the axial compression increases the shear capacity of the section, and vice versa. As a result, VGM
can impose significant variation in axial compressive force and increase the shear capacity of the
section in cases when PGA of this component arrives earlier than other PGAs (i.e., for negative
time lag). On contrary, when PGA of HGMs arrive early than that for VGM (i.e., positive time
lag), it can cause significant inelastic deformation of piers and result in higher curvature ductility
demand and reduced shear capacity of piers. It can further be elaborated using the results presented
in Fig. 6.19, which illustrates the shear capacity of a pier of the representative bridge under all six
ground motions under pristine and 100-year degraded state of the bridge. The maximum observed
CDD of the pier section under each seismic event is also shown in the figure for an easy reference.
The shear capacity of the section under seismic event with negative time lag (i.e., GM # 5) has the
highest capacity among all other records. This may be due to the exposure of early vertical ground
acceleration during the seismic event. In contrast, the seismic events with positive arrival time-
lags (i.e., GM # 1, 4, 6) have lower shear capacities than that for GM # 5. However, no direct
correlation between the arrival time lag and the shear capacity of piers can be established based on
limited observations. Additionally, curvature ductility of piers also plays crucial role here. Perhaps,
this is the reason of observing lower shear capacity of the pier under GM # 1 than that under GM
# 6, even if the former has lower arrival time lag than the later. In every scenario, though,
corrosion-fatigue degradation caused lower shear capacity of the pier for obvious reasons.

161
Figure 6.18 Curvature ductility demand of pier at pristine and 100-years of service under all
considered ground motions

Figure 6.19 Shear capacity of a bridge pier under six ground motions

162
6.7 Vulnerability analysis of corrosion-fatigued bridges under the combined impact of
VGM and HGM

As mentioned previously under research gap, no previous research thus far has been
conducted to establish the probabilistic seismic vulnerability of aging RC fatigued bridges under
the combined influence HGM and VGM. Therefore, the overall objective of this section is to
establish a seismic fragility model for each bridge component of RC bridges, taking into the
consideration of combined impact of HGM and VGM.

6.7.1 Selection of input ground motions to generate fragility curves

The selection of appropriate ground motion records is a crucial task in developing fragility
curves for highway bridges. Previous studies have demonstrated that the variability among ground
motion records is the primary contributor to the aleatory uncertainty in resulting fragility curves
(Jiang et al. 2021). As a result, the selection of ground motion is expected to significantly affect
the resulting fragility curves. Due to the availability of limited number of recorded ground motion
for Gujarat, it is dire necessary to explore additional source of data for the seismic input. In addition
to the six-ground motion which were initially selected based on the frequency content of 2001
Bhuj earthquake (refer Table 6.3), acceleration time-history that have been recorded in other parts
of the world, and are compatible with fault mechanism and seismic potential of Gujarat are being
additionally obtained to perform dynamic analysis, and this selection method is discussed in
Section 6.5.

In total, a set of 30 ground motion record are chosen from PEER NGA database (PEER
2023), and the list of selected ground motions are listed in Table C5 of Appendix C. Fig. 6.20a
shows the response spectra of 30 ground motion in the horizontal direction along with the design
(i.e., target) spectrum (Zone V, as per BIS (2002)), as well as the mean and dispersion spectra
which are bounded by one standard deviation in upper and lower limits. It can be observed that the
mean spectra of 30 selected ground motion matches well with the target spectrum and are bounded
within the upper and lower bound of dispersion spectra.

163
Figure 6.20 (a) Target and mean spectrum of selected ground motion from PEER NGA database
(b) V/H ratio of selected ground motions

The V/H ratios of these selected ground motions and that for the recorded 2001 Bhuj
earthquake in Gujarat are shown in Fig. 6.20b for the time period ranging between 0.01-10s. The
thick black line represents the two-third threshold specified in code provision, which is exceeded
by at least 50% of the dataset. Notably, this ratio is exceeded not only for earthquakes with M w >
6, but also for other records having Mw <6. Additionally, it is observed that the V/H ratio exceeding
two-third is restricted to a time period range of 0.01 to 5 sec, which coincides with the typical
fundamental time period of short-to-medium span bridges.

6.7.2 Intensity of selected ground motion

Different ground motion parameters can be used to represent the intensity measure (IM) of
ground motions, depending on specific response characteristics that govern the behaviour of the
structure. It was already observed in Section 6.6 that the seismic behaviour of RC bridges may get
altered due to corrosion fatigue degradation, particularly in the vertical mode of vibration. Hence,
it is crucial to account for various possible ground motion IMs and determine the most appropriate
one for assessing seismic vulnerability of bridges. Several studies have been conducted in this
regard to evaluate the suitability of IMs that are either specific or non-specific to structures under
consideration. Some of the identified non-structure- and structure-specific IMs are listed in Table
6.5 and Table 6.6. In a study conducted by Mackie and Stojadinovic (2001), the suitability of
fifteen IMs of ground motions for California bridges was judged and indicated that the most

164
suitable IMs are spectral acceleration and spectral displacement at the natural vibration period of
the bridge. According to Padgett et al. (2008), PGA is an appropriate ground motion IM
considering the sufficiency, efficiency, practicality, and the ability to quantify seismic hazard.
However, unlike spectral acceleration or displacement at the natural period of bridge structures,
PGA as the IM cannot directly correlate structural response with ground motions. Other existing
studies explored the appropriate selection of IMs which are subjected to near and far-fault ground
motions for bridges. For an instance, Wang et al. (2013) performed an investigation on highway
bridges by accounting for eighteen IMs related to both horizontal and vertical ground motions
(listed in Table 6.5 and Table 6.6), and identified PGV as the most suitable IM for the demand of
each component excited by HGM and VGM. Another study by Wang et al. (2018) analyzed a
bridge supported on pile shafts in soil prone to have liquefaction and lateral spreading. This
research indicated that velocity-based IM would be appropriate for developing probabilistic
seismic demand model (PSDM) when compared to other ground motion IMs that are based on
acceleration, displacement, and spectral period. In contrast, Zhong et al. (2019) indicated PGV as
an appropriate IM for a cable stayed bridge that is subjected to near and far-fault ground motions.
Dehghanpoor et al. (2021) studied seismic performance of isolated bridges and mentioned PGV,
Housner spectrum intensity, vertical spectral acceleration at time period 0.2 sec, and square-root-
of-the-sum-of-square (SRSS) of vertical spectral acceleration at first and second vertical period of
bridge structures to be the optimal IMs of ground motions. Eslamnia et al. (2023) utilized an
energy-based approach to determine the optimal IM among 24 selected parameters, and indicated
that the velocity spectrum intensity (VSI) associated with the horizontal ground motion (HGM)
serves as the optimal IM for characterizing seismic demand for highway bridges which are located
in near-fault regions. Altogether, previous studies have primarily indicated PGV as the ideal IM
for characterizing the seismic demand of bridges components that are excited by both HGM and
VGM, assuming a scalar IM derived solely from HGM. However, Wang et al. (2013) indicated an
alternative approach that considers EDP which are influenced by VGM, such as bending moment
of bridge deck and axial force on columns. The study indicated that the optimal IM for such
scenarios would be square root of sum of squares (SRSS) of vertical spectral accelerations at the
first and second vertical modes of bridge which are affected by VGM. The current study utilizes
the same IM as suggested in Wang et al. (2013) to demonstrate the lifetime vulnerability of a case-
study bridges that is influenced by corrosion-fatigue and VGM

165
Table 6.5 Non-structure specific IMs which are identified in literature

S. No IM Description Formulation Symbol


notation

𝑢̈ 𝑔 = ground
1 PGA Peak ground acceleration Max |𝑢̈ 𝑔 |
acceleration

𝑢̇ 𝑔 = ground
2 PGV Peak ground velocity Max |𝑢̇ 𝑔 |
velocity

𝑢𝑔 = ground
3 PGD Peak ground displacement Max |𝑢𝑔 |
displacement

1 2 g = acceleration
4 Arias Arias intensity ∫(𝑢̈ 𝑔 ) 𝑑𝑡
𝑔 due to gravity

T = total time
𝑇 2
1
5 RmsA Root means square acceleration √∫ (𝑢̈ ) 𝑑𝑡 duration of
0 𝑇 𝑔
ground motion

Root means square 𝑇 2


1
7 RmsD √∫ (𝑢̇ ) 𝑑𝑡 -
displacement 0 𝑇 𝑔

8 CAV Cumulative absolute velocity ∫|𝑢̈ 𝑔 (𝑡)| 𝑑𝑡 -


0

𝑇
Cumulative absolute
9 CAD ∫|𝑢̇ 𝑔 (𝑡)| 𝑑𝑡 -
displacement
0

166
S. No IM Description Formulation Symbol
notation

Third highest
absolute
Sustained maximum
10 SMA acceleration -
acceleration
magnitude in
ground motion

𝑇
2
11 SED Specific energy density ∫|𝑢̇ 𝑔 (𝑡)| 𝑑𝑡 -
0

Table 6.6 Structure specific IMs which are identified in literature

S. No IM Description Formulation Symbol


notation

Spectral acceleration at a
1 Sa (T) - −
natural time period T of bridge

1
𝑛 𝑛
𝑛 = number of
2 Sa,avg Average spectral acceleration (∏ 𝑆𝑎 (𝑇𝑖 )) vibration mode
𝑖=1 of bridge

0.5

Acceleration spectrum ∫ 𝑆𝑎 (𝜁 𝑢𝑔 = ground


3 ASI
intensity 0.1 displacement
= 5%. 𝑇)𝑑𝑡

2.5
g=
∫ 𝑆𝑣 (𝜁
4 VSI Velocity spectrum intensity acceleration
0.1
due to gravity
= 5%. 𝑇)𝑑𝑡

167
S. No IM Description Formulation Symbol
notation

2.5
𝑃𝑆𝑣 = peusodo
∫ 𝑃𝑆𝑣 (𝜁
5 HSI Housner spectrum intensity spectral
0.1
velocity
= 5%. 𝑇)𝑑𝑡

6.7.3 EDP for the fragility analysis of bridge which are affected by corrosion-fatigue and
VGM

The literature had utilized various parameters to define and categorize the damage states
of RC pier under earthquake loading. These parameters include drift ratio, displacement, and
curvature ductility. As the adopted case-study bridge structure is anticipated to fail due to flexure
for its given dimension and height, the current study utilizes the curvature ductility of bridge
column to classify damage states for the input ground motions. The curvature ductility demand for
each seismic event is obtained by dividing the maximum curvature of the pier by the yield
curvature or the curvature at which the pier reaches its yield strength. Additionally, this process
accounts for the change in yield curvature of the section for different axial load ratios and also for
each life cycle year (as shown in Fig. 6.17a). In order to classify the obtained response into various
damage states, the demand is compared with the threshold limit as mentioned in Table 2.4 in
Chapter 2. Similarly, the moment demand for each girder in all span are obtained and compared
with its capacity for each life year of the bridge, and this enables to classify each girder as either
collapsed or not. The damage states for elastomeric bearing and abutment are determined and
classified according to the EDP mentioned in Table 2.2 and Table 2.3 in Chapter 2.

6.7.4 Component level fragility curves of case-study bridge

Fragility curves of abutments, elastomeric bearings, girders, and piers at their pristine
states, accounting for both HGM and VGM in the dynamic analysis for various damage states are
illustrated in Fig. 6.21. These curves are plotted based on the discussion presented in Section 2.3
in Chapter 2. The IM of the fragility curves are based on the sum-of-square-root (SRSS) of the
spectral acceleration of VGM at first two and second time period of vertical mode of vibration

168
(Sa12-SRSS). The fragility median of all curves is mentioned within figure. However, as discussed in
Section 6.6.1, corrosion and corrosion-fatigue degradation can alter the dynamic characteristic of
the bridge. Therefore, the structure specific selected IM is subjected to alter for different lifecycle
year of bridge and Table C6 in Appendix C presents the Sa12-SRSS for each selected ground motion
at different life-cycle years of the bridge.

From Fig. 6.21, it can be observed that at pristine state, the pier is the only component that
is susceptible to major and collapse damage states, whereas the damage to elastomeric bearing is
limited to minor and moderate states. Furthermore, the abutment experience only minor damage
state, and the girder never reaches its collapse state for the given input of ground motions. Fig.
6.22 also presents the component-level fragility curves of a case-study bridge at its 75th and 100th
year of service life. These curves are developed by taking into account the residual area of
embedded steel in the girder caused by corrosion-fatigue, based on a fatigue stress of 7.5 MPa
(refer Fig. 6.10a). At both 75 and 100 years of service, it can be observed that the fragility median
of abutment is not subjected to change due to the presence of damage in girders and piers.
However, the fragility median of elastomeric bearing has shown an increase in vulnerability at the
75th year of service life, with 13% and 28% increase with relative to its pristine state at minor and
moderate damage state, respectively. Similarly, at 100-year of service life, the vulnerability of the
elastomeric bearing has increased by 13% and 31.7% relatively to its pristine state. The moment
demand on the girder from the seismic excitation never exceeded its capacity at its pristine state
or at its other-life cycle year. Therefore, the probability of failure for the girder under given seismic
event is essentially zero by the 75th year of service life. Nevertheless, it was discussed earlier that
flexural capacity of the girder decreases due to rebar damage caused by corrosion-fatigue (refer
Section 6.6.2). As a result, the girder cannot bear its moment demand beyond its capacity after 100
years of service, and results in collapse, and the corresponding fragility median is observed to be
0.85g. For the extent of corrosion damage to the rebar in pier, it was determined that its
vulnerability of bridge at 75-year service life had increased by 9.7%, 12.1%, 11.6% and 11.6% for
its minor, moderate, major and collapse damage states, respectively. Subsequently, at 100-year
service life, the pier vulnerability increased by 12.4% for the collapse damage state, while the
percentage changes for the other damage states remained the same as 75-year service life.
Additionally, it has to be noted that the seismic vulnerability of the all components at 25 and 50
years of service is same as it was in pristine state. Overall, these figures provide valuable

169
information into the seismic behavior and performance of different bridge components under
different damage conditions, and this can aid for informed design and maintenance decision to
improve the seismic safety of bridges.

Figure 6.21 Fragility curves for components such as (a) abutment (b) bearing (c) girder (d) pier –
in their pristine state and at different levels of damage.

170
Figure 6.22 Fragility curves for components such as (a) abutment (b) bearing (c) girder (d) pier –
in their 75th and 100 year of service life and at different levels of damage.

171
6.7.5 System level fragility curves of case-study bridge

In order to determine the seismic vulnerability of the bridge system, an overall evaluation
must be performed by taking into account the impact of all bridge components. Fig. 6.23 presents
the system level fragility curves of a case-study bridge, illustrating its seismic vulnerability at
different damage state and as well as at different life-cycle year. It can be observed that up to 50
years of service life, there was no observable change in the seismic vulnerability of bridge at minor
damage state. However, at 75th year of service, the seismic vulnerability of bridge has increased
by 5% relative to that in the pristine state. Table 6.7 presents the fragility median for the bridge
system and along with its constituent components. The data indicates that the fragility median of
elastomeric bearing at minor damage state is 0.23g, which is 21% higher than the system-level
median. However, the fragility median for the abutment and pier exhibit difference of over 50%
as compared to the system-level median, highlighting the partial contribution of both elastomeric
bearing and pier to the system level fragility in the minor damage state. The moderate damage
state of the case-study bridge has experienced an increase of 28% at 75th year of its service life.
From Table 6.7, it can be observed that the fragility median of elastomeric bearing at moderate
damage state is 0.59g, which is equivalent to the system-level median. This indicates that the
elastomeric bearing plays a crucial role in determining and dominating the moderate damage state
of the bridge at system level. It is observed that the fragility median of the bridge in major and
collapse damage state is 2.34g, which is the same as the median observed for piers.

This suggest that the seismic performance of pier plays a dominant role in determining the
extent of damage, particularly in cases of major and collapsed damage state. At 100 years of
service, the bridge shows a similar pattern of damage as such as in 75 years of service. Pier and
elastomeric bearing are partially responsible for causing minor damage to the bridge system, while
the moderate damage is influenced mainly by the seismic performance of elastomeric bearing.
However, at 100 years of service, the seismic performance of girder primarily governs the major
and collapse damage state of bridge.

172
Figure 6.23 Time-dependent system level fragility curves at (a) minor (b) moderate (c) major (d)
collapse damage state

173
Table 6.7 Computed median fragility values of the bridge system and its associated components for the 75th year service life

System Abutment Elastomeric bearing Pier Girder


Damage
state Median Median Difference Median Difference Median Difference Median Difference
(g) (g) (%) (g) (%) (g) (%) (g) (%)

Minor 0.19 2.65 +1295 0.23 +21 0.28 +52.6 - -

Moderate 0.59 - - 0.59 0 1.23 +108.5 - -

Major 2.34 - - - - 2.34 0 - -

Collapse 2.34 - - - - 2.34 0 - -

Table 6.8 Computed median fragility values of the bridge system and its associated components for the 100th year service life

System Abutment Elastomeric bearing Pier Girder


Damage
state Median Median Difference Median Difference Median Difference Median Difference
(g) (g) (%) (g) (%) (g) (%) (g) (%)

Minor 0.19 2.65 +1372.2 0.23 +27.7 0.28 +52.6 - -

Moderate 0.48 - - 0.59 +16.7 1.23 +108.5 - -

Major 0.85 - - - - 2.34 0 - -

Collapse 0.85 - - - - 2.34 0 0.85 0

174
6.8 Closure

This chapter presents an improved model for predicting the deterioration of rebar in bridges
due to corrosion and fatigue. The model considers the effect of both pitting corrosion and fatigue
cracks induced by the passage of truck traffic on the degradation of embedded rebar in girders.
Relevant data on corrosion and traffic are obtained and used to accurately model the degradation
process over the lifespan of the bridge. The residual area of rebar is calculated for different level
of fatigue stress, and the results show a significant reduction after 100 years of service. The model
is used to assess the change in flexural stiffness of the bridge deck and the seismic vulnerability of
the bridge under different stages of service life.

A case study to demonstrate the application of developed framework is performed on a


representative bridge in Gujarat, India. The study investigates how the modal parameters of the
bridge change with deterioration owing to corrosion-fatigue and corrosion in girders and piers.
The result indicates the deterioration of the bridge deck significantly reduces the frequency of
vertical model of vibration, which could potentially amplify the response of bridge components
during a seismic event. The study explores the impact of deck deterioration on the moment demand
and capacity of the girder, longitudinal deformation of elastomeric bearing, and curvature ductility
demand and shear capacity of pier. Time-dependent seismic fragility model for each component
of RC bridge was developed taking into the account of the combined influence of HGM and VGM
in the dynamic analysis. The study indicated that there was no change in the seismic vulnerability
of bridge components within the first 50 years of its service life. The seismic vulnerability of the
elastomeric bearing and pier increased from the 75th year onwards due to deterioration of the girder
and pier. The elastomeric bearing and pier are found to dominate the vulnerability of entire bridge
system. At the end of bridge service life, in addition to pier and elastomeric bearing, the seismic
performance of girder dominates the higher damage states such as major and collapse. The results
indicate that neglecting the corrosion fatigue mechanism while using the conventional
deterioration model would result in an underestimated system performance and lower estimation
of its seismic vulnerability.

175
This page is intentionally left blank

176
CHAPTER 7

SUMMARY AND CONCLUSION

This chapter provides a summary of the research conducted as part of this dissertation. This
includes an overview of major findings, contributions made by this research, design
recommendations for standard provisions, and the potential directions for future work.

7.1 Summary

This research aimed to develop a quantitative framework for assessing the life-cycle multi-
hazard resilience of RC bridges which are spanning across rivers and situated in urban areas. The
interaction of earthquake and flood results in the formation of multi-hazard scenario for bridges in
seismically active and flood-prone areas. Bridges located in seismically active urban areas are also
vulnerable to multi-hazard resulting from the interaction of environmental corrosion and
mechanical fatigue. A quantitative framework has been established for assessing time-dependent
seismic vulnerability of bridges, considering the hazard interaction which changes the condition
of bridges. For an example, flood results in scouring of soil sediment around piers, while corrosion
and corrosion-fatigue result in dissolution of efficient cross-sectional area of embedded rebars in
piers and girders. Both of these conditions increase the flexibility of bridge systems and make them
more vulnerable for seismic hazard.

Over time, climate change affects the safety and stability of riverine bridges by altering
flood patterns. This concern was addressed in the study by developing an integrated approach to
evaluate the potential impact of climate change on multi-hazard performance involving earthquake
and flood. The application of this framework is demonstrated on an existing bridge over the San
Joaquin River in California. Based on the observed streamflow in the river between 1930 to 2011
and applying climate change projection model for the region, projected flow in the river for the
two consecutive projection periods between 2012 to 2050 and 2051 to 2099 are developed. The
observed and projected data are used to generate flood hazard curves and corresponding scour
depths for 100-year floods. By analyzing these scoured bridges under seismic motions, it is

177
observed that climate-change induced scour had a greater influence on the seismic performance of
piers when compared to other components of the bridge (such as abutment and elastomeric
bearings). Hence, the influence of various bridge design parameters for piers (such as aspect ratio
and longitudinal reinforcement ratio of piers) and the differential ground elevation between pier
bents is studied to evaluate the multi-hazard performance of river-crossing bridges. It was found
that the aspect ratio and longitudinal reinforcement ratio of piers can significantly influence the
multi-hazard performance of such bridges. The findings also demonstrated how design parameters
may be revised to perform risk-targeted multi-hazard design of bridges.

Bridges located in urban areas are prone to experience fatigue due to evolution and growth
of truck traffic. This scenario, coupled with environmental corrosion, can lead to the condition
referred as corrosion-fatigue degradation, which affects the structural performance of bridge decks.
The study proposed a quantitative framework to quantify the residual cross-sectional area of
embedded rebars in girders, taking into account of corrosion-fatigue degradation. The quantified
residual area is further incorporated into the FEM of a representative bridge, and NLTHA are
performed for two bridge models having two and three identical simply supported spans to
evaluate their seismic vulnerability at different stages of bridge service life (i.e., at intact, 25, 50,
75, and 100 years of deterioration). The result provided valuable insights into the effect of
corrosion-fatigue on the structural integrity of bridges. The study also found that corrosion-fatigue
has a negative impact on the flexural stiffness of girders. As a result, the exposed girder is found
to be weaker when subjected to earthquake loads, especially those with three significant
translational components.

Further, the study illustrated the significant impact of corrosion-fatigue and VGM on the
seismic response of different bridge components, such as elastomeric bearing, girders and piers by
analysing a representative multi-span RC bridge located in Gujarat, India. Initially, the bridge is
analysed for six ground motions that represent recorded seismic ground motions in the Gujarat
region. Research outcomes suggest that the combined impact of the corrosion-fatigue degradation
and VGM can lead to higher seismic vulnerability of bridges and endanger their safety under future
earthquakes. Additional NLTHA are also performed on the bridge model, and time-dependent
seismic fragility model for each component of RC bridge was developed taking into the account
of the combined influence of HGM and VGM. The result indicates that neglecting the corrosion-

178
fatigue mechanism while using the conventional deterioration model would result in an
underestimated system performance and a lower estimation of failure probability of bridges.

7.2 Conclusion

Important conclusions drawn from this research are listed below

1. Projected intense flood hazard due to regional climate change increases the exposure of the
bridge to flood-induced scour around bridge piers beyond design limits. It leads to an
increased maximum curvature ductility demand of piers under seismic events, indicating a
rising trend in bridge seismic response. Even if there is no notable variation in seismic
vulnerability of the bridge at the minor and moderate damage states, the same is quite
significant at the major damage and collapse states. At the major damage state, fragility
medians of the bridge are observed to increase by 11% and 17.3% for flood scenario II and
III, respectively, from that for scenario I. These values at the collapse state are 15.1% and
21.2%.
2. The increase in flood flow owing to climate change also increases the seismic risk of the
bridge at specific hazard level. As estimated, the potential annual loss due to the damage
of the bridge under future seismic activities increases from USD 7.6 million (for flood
scenario I) to USD 8.7 million (for flood scenario II) and USD 9.2 million (for flood
scenario III). Consequently, seismic resilience of the bridge drops by approximately 14%
at a particular seismic hazard level. If single hazard events (such as flood-induced scour
alone) were to be considered, the bridge might not have experienced such significantly
larger risk due to climate change projection.
3. An adaptation plan is implemented to decrease the expected risk and to enhance seismic
resilience of the investigated bridge by incorporating rip-rap confinement around bridge
piers. A cost-benefit analysis of the adaptation plan demonstrated its favorable socio-
economic impact over the bridge remaining service life. Such analyses provide insight into
policy making for prioritizing and implementing climate change adaptation plans for
bridges.
4. For bridges under the multi-hazard condition of earthquake and flood-induced scour, pier
aspect ratio is found to be critical for bridge performance under both seismic only and
multi-hazard conditions. 10% higher pier diameter helped to increase the moment capacity

179
and flexural stiffness of piers. It resulted in lowering bridge seismic vulnerability and risk
of bridges with and without scour. On contrary, moment capacity and flexural stiffness of
piers got increased for 10% decrease in pier diameter, and thus, resulted in higher
vulnerability and risk of bridges than that of the base models.
5. The variation in longitudinal reinforcement ratio in pier is impactful for river-crossing
bridges subjected to multi-hazard condition. A 10% reduction in longitudinal
reinforcement ratio without changing the pier diameter results in reduction in the moment-
carrying capacity and ductility of the piers, which in turn produced higher multi-hazard
vulnerability and risk of scoured bridges. Differential riverbed elevation at pier bases did
not produce any significant impact. Overall, the study suggests revising bridge design to
endure targeted risk performance level under specified hazard conditions.
6. For bridges that are susceptible to corrosion-fatigue degradation, the residual area of
embedded rebars in girder is observed to reduce to 49% and 70% (mean values),
respectively, for aggressive and conservative traffic growth models. This reduction can be
higher or lower if uncertainties from various sources of input are considered. In addition to
damage in rebar, the fatigue also reduces the modulus of elasticity of concrete. The product
of two degradations (i.e., modulus of concrete and moment of inertial of girder) results in
reduction of flexural stiffness of bridge decks.
7. Results from NLTHA of such corrosion-fatigued bridges depict an increase in seismic
vulnerability of the bridges that can be measured by comparing the median value from
corresponding fragility curves. Estimated values reveal that the three-span bridge has
42.5%, 56.7%, and 43.8% increase in seismic vulnerability at the end of 100-year life,
respectively, at moderate, major damage and collapse states. Similar observations are made
for the two-span bridge.
8. Due to the increase in bridge seismic vulnerability caused by corrosion-fatigue
degradation, both two- and three-span bridges exhibit significant reductions in resilience
values at various stages along their life cycles. The two-span bridge experiences maximum
reductions of 10.4% and 13.1% in seismic resilience at 75th and 100th year of its service
life, respectively, while the corresponding reductions for the three-span bridge are 11.1%
and 14.1%, respectively. These findings indicate that corrosion-fatigue degradation
significantly reduces the seismic resilience of the investigated bridges. It has to noted,

180
however, that observed results are not generally applicable to other bridges and/or other
exposure conditions.
9. Furthermore, it is observed that the reduction in residual area resulting from corrosion-
fatigue degradation is not only solely dependent on the type of traffic growth model, but
also on the magnitude of fatigue stress. Various magnitudes of fatigue stress such as 7.5
MPa, 4.5 MPa and 2.4 MPa are observed to result in, respectively, 46%, 25%, and 9%
reductions of residual area of rebars at the end of bridge service life.
10. The dynamic properties of the bridge are affected by the reduction in flexural stiffness of
constituent girders and piers. The alteration in dynamic properties of the degraded bridge
is most prominent for the vertical mode of vibration of the bridge, at which 35% reduction
in modal period is observed at the end of 100-year service life. Deteriorated girders
experience large amplification of moment demand from seismic events having significant
VGM component, leading to deck failure due to flexure at 100-year degraded state under
earthquakes with V/H ≥ 0.66.
11. The combined effect of corrosion-fatigue degradation and VGM increases the seismic
vulnerability of elastomeric bearings and piers. This results in higher translational
displacement of bearings and altered friction coefficient between girder surface and rubber
pads. Under the influence of VGM, the axial force on pier changes, resulting in a negative
impact on both curvature ductility demand (CDD) and shear capacity of pier. As pier CDD
increases, shear capacity decreases resulting it more vulnerable to earthquakes with VGM
components. Such increase in CDD and decrease in shear capacity may have some
correlation with the arrival time lag of seismic components. However, no definite
correlation can be established with the limited observations made in this research.
12. A time-dependent seismic fragility model was developed for each component of RC bridge
using ground motions with significant HGM and VGM components. The study indicated
that the seismic vulnerability of elastomeric bearings and piers decreased after 75 years of
service life due to the deterioration in girder and pier. At the pristine state, seismic
performance of piers and elastomeric bearings govern (partially) the minor damage of the
bridge, while moderate damage is mainly influenced by the seismic performance of
elastomeric bearings. Higher damage states such as major and collapse are governed by the
seismic performance of piers. At 100 year service life of the bridge, however, the seismic

181
performance of girders primarily governs the major and collapse state of the bridge. This
observation signifies that neglecting corrosion-fatigue while using the conventional
deterioration model would result in an underestimated system performance and lowered
failure probability of bridges.

7.3 Key research contribution

This research contributes to two broad areas of time-dependent probabilistic seismic


vulnerability assessment of RC highway bridges that are situated across river and at urban areas
with dense truck traffic. As a result, the study has made significant contributions to these two
specific areas:

I. In the context of river-crossing RC highway bridges

1. The research has enhanced the existing methodology for seismic performance
evaluation of riverine bridges by incorporating climate change as a geoclimatic stressor
in assessing the seismic losses and resilience of bridges in the long run. This is the first-
of-its-kind study to assess multihazard risk and resilience of riverine bridges under
global warming and climate change.
2. The study also contributed new knowledge on the impact of specific bridge design
parameters on the multi-hazard performance of riverine bridges, providing insight into
how to revise bridge design for targeted outcome. By understanding the individual
impact of design parameters related to bridge piers, informed decisions can be made to
revise bridge design and achieve specific targeted risk performance.
II. In the context of RC highway bridges at urban areas with dense truck traffic

1. The study introduced a novel method for assessing seismic vulnerability of RC


highway bridges, which considers the coupled interaction of environmental
corrosion and traffic fatigue on the gradual degradation of embedded rebar in
girders. The approach integrates various modulus for analysing stochastic traffic
(based on limited observations), vehicle bridge interaction to generate stress in
girders due to the passage of trucks, and material degradation owing to corrosion-
fatigue phenomenon. Thus, the study provides a way to eliminate the need for in-
situ test to measure fatigue-induced stresses in bridge girders.

182
The research generated in-depth knowledge on the effect of VGM on life-cycle seismic
response of bridges subjected to corrosion-fatigue degradation. Thus, it improves the currently
available numerical model for assessing the remaining cross-sectional area of embedded rebars in
girders that undergo corrosion-fatigue damage over time. The updated model incorporates the
pitting corrosion mechanism and fatigue cracks to quantify the residual area of embedded rebars,
resulting in a more accurate and comprehensive evaluation of the residual cross-sectional area of
rebars subjected to corrosion-fatigue degradation.

7.4 Design recommendation for bridges in India

1. The current design guidelines of Indian Road Congress (IRC) of concrete bridges such as
IRC:6-2017, IRC: 112-2020 do not explicitly account for multi-hazard effects into the
design of bridges. To ensure a holistic approach to multi-hazard design of bridge, the
research outcome recommends to consider, prior to bridge design, the correlation and
dependency of various regional hazards that may pose a threat on safety and serviceability
of bridges in the long run.
2. To incorporate the effect of climate change in scour calculation for bridge design, it is
recommended to consider the potential increase in flood magnitude and frequency due to
climate change. This could be done by performing flood-frequency analysis on the
historical flood data and projecting future changes in flood patterns based on available
appropriate climate models. The resulting estimate of flood level should be used to
evaluate hydrodynamic and water current force. Moreover, as mentioned in the IRC:6-
2017, the depth of scour under seismic condition is considered to be 0.9 times the
maximum scour depth. It is suggested to reconsider the value of 0.9 in the context of
potential changes in sediment transport due to regional climate change.
3. It is widely recognized that material properties of embedded steel rebars (that are used
while designing bridges) are subject to change over time due to corrosion and other
deterioration mechanism. These mechanisms can significantly affect the structural
performance throughout its service life. To address this issue, it is essential to incorporate
load and resistance factors that account for degradation caused by aging, corrosion,
corrosion-fatigue and carbonation during the service life. Moreover, the factor of safety,
which is been evaluated at the design stage of bridges, should be evaluated throughout the

183
entire service life of bridges to ensure the desired performance and safety of bridges during
lifetime. Therefore, it is crucial to adopt an integrated approach that accounts for life-cycle
based load and resistance factors for the analysis and design of bridge structures.

7.5 Limitation and future work

Although this dissertation has made several contributions in advancing the existing
knowledge for assessing multi-hazard resilience of RC highway bridges, there are some limitations
of the current research that can be explored more in future. These future research areas are listed
below.

1. The current approach assumes that the use of riprap around bridge piers completely restores
bridge flexibility and brings bridges to their pre-scour conditions. It is also assumed that
no additional scouring will occur in future due to the presence of ripraps. To justify these
assumptions, further investigations are needed to determine the extent to which riprap can
restore lost flexibility of bridge.
2. The seismic performance of piers can vary depending on their shapes and the boundary
condition of bridge decks (whether continuous or simply supported). The study in this
dissertation focused on a circular-shaped piers; therefore, further research is necessary to
investigate the seismic performance of other pier shapes and superstructure types. As a
result, the findings presented in Chapter 4 should not be applied to all RC highway bridges.
3. The formation of rust cue to corrosion expands inside the concrete and lead to the formation
of cracks in cover concrete. The numerical model developed to calculate the residual area
of embedded rebars due to corrosion-fatigue degradation did not consider the additional
penetration of chloride ions through these cracks. Therefore, future research should be
performed to modify the numerical model to investigate the influence of cracks or early-
age cracks on residual area of rebars under corrosion-fatigue degradation.

Finally, the VBI analysis considered bridge girders to remain elastic. This assumption can
be updated by considering the nonlinear behaviour of girders under higher traffic loads.

184
This page is intentionally left blank

185
APPENDIX - A
Synthesis of studies that assessed the performance of highway bridges under multi-hazard scenario

This section in Appendix A documents a record of various studies which are conducted until 2018 to evaluate the seismic
performance of bridges in the face of multi-hazard condition resulting from the interaction of two or more natural event.

Table A.1 Synthesis of studies (up to 2018) on bridge performance assessment under multi-hazard condition

Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Impact of high-rise pile


Scour has detrimental impact on
foundation with pile cap
Han et al. 2010 __ bridge seismic performance when __
on bridge seismic
river bed is much lower than pile cap
vulnerability

Flood discharge,
Scour and Seismic fragility curves Seismic fragility and risk of bridges
Banerjee and PGA and Scour calculation
Earthquake and risk in the presence increase in the presence of flood-
Prasad 2013 scour depth parameters, Diameter
and absence of scour induced scour
of piers

Seismic fragility curves Bigger diameter piles help to reduce


Prasad and PGA and
for various diameter of the impact of flood-induced scour on __
Banerjee 2013 scour depth
scoured piles bridge seismic fragility

186
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Field test on an existing No significant change in seismic


Chang et al. PGA and
bridge with scoured performance of the bridge with scour __
2014 scour depth
caisson foundation depth upto 4m

Impact of scour on
PGA and
Wang et al. seismic fragility curves Scour can result in lower collapse risk
SAs at 2, 5 __
2014a of bridges on pile for piers, but higher risk for piles
and 10 secs
foundations

Calibration of load PGA and Flood discharge,


Wang et al. Proposed scour load factors for short-
factors for multihazard SAs at 2, 5 Scour calculation
2014b and medium-span bridges
bridge design and 10 secs parameters

Experiments on bridges With increasing depth of scour,


Wang et al. 2015 with scoured piles under PGA bending moment in piles increases and
seismic excitations the same in piers decreases

Chandrasekaran Optimum retrofit Composite jackets are


and Banerjee strategies for resilience PGA observed to improve seismic resilience __
2016 enhancement of bridges with scoured foundations

Dong and Seismic fragility and PGA and Seismic vulnerability and expected Flow intensity and
Frangopol 2016 functionality of bridges scour depth loss increases with scour occurrence interval

187
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

in the presence of flood-


induced scour

Multihazard fragility PGA and Significance of flood hazard on bridge


Gehl and
surfaces using Bayesian flood seismic fragility is observed at __
D’Ayala 2016
Networks discharge collapse state of the bridge

Impact of scour and Scour can enhance annual and


Guo and Chen Scour depth
earthquake in lifecycle cumulative probabilities of extensive __
2016 and
of bridges damage and collapse of bridges

Multihazard fragility Large diameter pile shafts minimize


Yilmaz et al. curves, surfaces and risk PGA and scour impact on bridge seismic
__
2016 of two real-life bridges scour depth vulnerability, whereas traditional pile
in California foundations may increase the same

Uncertainties in fragility Key uncertain parameters are


Parameters related to
Yilmaz et al. curves and risk of a real- PGA and identified; Possible variations in
bridge structure and
2018 life bridges under scour depth multihazard fragility and risk of the
underlying soil
regional hazards bridge are evaluated

188
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Flood-induced scour can posse


Impact spectrum of
Yilmaz and PGA and positive and negative impacts on
scour on bridge seismic __
Banerjee 2018 scour depth seismic vulnerability of bridge
performance
components

Seismic fragility curves


Corrosion in steel rebars resulted in Corrosion process
Choe et al. 2009 of bridges with corroded SA
higher seismic vulnerability of bridges parameters
piers

Cumulative seismic Ground motion and


Corrosion in steel rebars enhances
Kumar et al. damage and life-cycle corrosion process
__ fragility and life-cycle cost due to
2009 cost of bridges with parameters
Earthquake and earthquakes
corroded piers
corrosion
Choe et al. 2010; Seismic fragility Increment functions can provide
Gardoni and increment function of __ fragility at any deteriorated stage __
Rosowsky 2011 corroded bridge piers based on initial fragility condition

Seismic fragility curves Seismic vulnerability of bridges


Ghosh and
of bridges with corroded PGA increases with time; Time-variant Corrosion parameters
Padgett 2010
piers and bearings fragility parameters are calculated

189
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Loss of bond between concrete and


Seismic fragility
SA in both surrounding steel was assumed
Simon et al. contours of a bridge
translational negligible; hence, no significant effect __
2010 with corroded pier and
direction of corrosion in pier and pile on bridge
pile
seismic fragility was observed

Beyond corrosion initiation,


Cumulative failure
cumulative failure probability is
Akiyama et al. probability with time of
PGV higher in case of earthquake and __
2011 corroded bridge piers
corrosion than that is for without
under seismic loading
corrosion

Seismic fragility and


Seismic fragility degrades with time
Alipour et al. life-cycle cost analyses
PGA due to corrosion and failure cost __
2014 of bridges with corroded
increases
piers

Expected seismic loss increases due to


Seismic loss assessment
Ghosh and corrosion; losses due to probable Loss estimation
of bridges with corroded PGA
Padgett 2011 seismic damage of various bridge model parameters
piers and bearings
components are calculated

190
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Neutralization (carbonation) of
Seismic fragility of concrete is done to simulate corrosion
Sung and Su
bridges with corroded PGA in concrete; seismic fragility degrades __
2011
piers and retrofitting cost increases with
time due to bridge neutralization

Effect of cracking in concrete cover


Seismic fragility of on rebar degradation due to corrosion
Zhong et al. PGA, PGV, Corrosion process
bridges with corroded is studied; no significant change in
2012 SA parameters
piers seismic fragility if less corrosion
damage in rebars

Seismic fragility of an
Choine et al. Bridge deck has negligible Corrosion process
integral bridge with PGA
2013 contribution to the damage parameters
corroded piers

Risk of aging bridges Seismic risk of bridges in the presence Traffic and
Decò and
due to seismic and __ of traffic loading increases along consequence analysis
Frangopol 2013
traffic loading bridge service lives parameters

191
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Transverse reinforcements may get


Seismic performance of
corrosion earlier than longitudinal Corrosion process
Ou et al. 2013 bridges with corroded PGA
reinforcements in piers depending on parameters
piers
bridge exposure condition

Prioritizing seismic
Deteriorated bridges in a network are
Rokneddin et al. retrofit of bridges in a Corrosion and
PGA ranked according to their criticality for
2013 network exposed to structural parameters
performing seismic retrofitting
corrosion

Corrosion in piers is visualized


through X-rays. Cumulative failure
Akiyama and Seismic performance of probability with time may increase
PGV __
Frangopol 2014 corroded bridge piers rapidly after corrosion initiates in case
of earthquake and corrosion than that
is for without corrosion

Seismic performance of Seismic performance of bridges with


Biondini et al. Corrosion and
bridges with corroded PGA corroded piers degrades with time;
2014 structural parameters
piers hence, design should consider this

192
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

Ghosh et al. Fragility and reliability Proposed bridge reliability in network


Structural, corrosion
2014 of a bridge network (BRAN) methodology to assess
PGA and bridge
Rokneddin et al. based on a limited set of reliability of bridge networksbased on
parameters
2014 instrumented bridges a limited set of instrumented bridges

Resilience degrades with bridge life


Biondini et al. Seismic resilience of
PGA due to gradual degradation from Corrosion process
2015 corroded bridges
corrosion

Seismic functionality of
Progressive deterioration of bridges
Alipour and a highway transportation
PGA results in higher seismic loss in the Corrosion process
Shafei 2016 network consisting of
network
corroded bridges

Seismic fragility of
Ni Choine et al. Corrosion may result in significant
bridges with corroded PGA Structural parameters
2016 reduction in bridge seismic fragility
piers

Dong and Resilience of corroded Expected seismic resilience of bridges


PGA __
Frangopol 2016 bridges over life-spans degrades due to corrosion degradation

193
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

An improved corrosion Seismic fragility degrades due to


Corrosion process
Ghosh and Sood model and seismic corrosion; concrete cover and water-
PGA and structural
2016 fragility of bridges with cement ration are observed to have
parameters
corroded piers impacts on corrosion effects on piers

Parameters in hazards
Effects of inspection frequency and
Life-cycle seismic associated with
Thanapol et al. results and hazards associated with
reliability of bridge piers PGV airborne chloride and
2016 airborne chloride on seismic failure
in marine environment steel weight loss
probability of a bridge are discussed
estimation process

Seismic fragility
Corrosion results in degraded seismic Corrosion process
Rao et al. 2017 function of corroded PGA
fragility of piers parameters
bridge piers

Seismic fragility of
Pitting is more severe than uniform
bridges and life-cycle
corrosion for bridge seismic Corrosion process
Shekhar et al. cost for various
PGA performance; bridge exposure and structural
2018 corrosion exposure
scenarios, according to severity on parameters
conditions and corrosion
seismic performance, are marine
processes

194
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

atmospheric zone, marine splash zone


and deicing salt spray

Bridge seismic fragility


Alipour et al. PGA and Calibrated scour load modification
curves in the presence of __
2013 scour depth factor for seismic design of bridges
scour

Earthquake, Corrosion
Seismic fragility curves PGA and Scour and corrosion enhance seismic
Scour and Dong et al. 2013 parameters, traffic
with scour and corrosion scour depth vulnerability and associated risk
Corrosion parameters

Seismic fragility in the


PGA and Increased seismic vulnerability in the
Guo et al. 2016 presence of corrosion __
scour depth presence of scour in corroded bridges
and scour

Risk of deteriorated Individual risk of deteriorated bridges


Earthquake, Corrosion process,
Decò and bridges due to traffic Spectral due to traffic load, flood and
scour, traffic structural and traffic
Frangopol 2011 load, flood and acceleration earthquake are estimated and then
and corrosion parameters
earthquake combined to obtain total risk

Earthquake Geotechnical and


Zhang et al. Fragility functions of Structural characterization has strong
and PGA, PGD bridge substructure
2008 various bridges due to roll in fragility of bridges under
liquefaction parameters

195
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

seismic shaking and seismic shaking and liquefaction-


lateral spreading induced lateral spreading

Soil liquefaction increases failure


Seismic fragility of steel Structural and
Aygün et al. probability of critical bridge
bridges on liquefiable PGA geotechnical
2011 components and uncertainty in
soils parameters
component fragility curves;

Seismic fragility of Parameters related to bridge structures


Structural and
Padgett et al. critical components of and underneath liquefiable soil can
PGA geotechnical
2013 bridges located on introduce a wide variation in bridge
parameters
liquefiable soil seismic fragility curves

Bridges performance Vertical component of ground motion


under horizontal and has great influence on responses of
vertical ground motion bridge piers; liquefaction may have
Wang et al. 2013 SA __
components in beneficial or detrimental influence on
liquefiable and non- the vertical seismic response
liquefiable soils

Wang et al. Seismic performance of Impact of seismic isolation is more


PGV __
2014c a seismically isolated pronounced in stiff soil than soft (or,

196
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

bridge on liquefiable liquefiable) soil, even if isolated


and stiff soils bridges on liquefiable soils perform
better than non-isolated bridges

Seismic performance of a RC duct


Seismic behaviour of a
Moshirabadi et located close to a bridge pier on
bridge pier on PGA __
al. 2015 liquefiable soil depends greatly on the
liquefiable soil
presence of pier

Understanding the Annual frequency of collapse of


Franchin and ability of a mainshock- mainshock-damaged bridge under
SA __
Multiple Pinto 2009 damaged bridge to carry aftershock decreases with time elapsed

earthquake traffic between the shocks.

sequence Following a mainshock, risk of a


Risk of mainshock-
(mainshock Alessandri et al. bridge is calculated for aftershock
damaged bridge to PGA __
and 2013 sequence to make a decision of bridge
aftershocks
aftershocks) usability after the mainshock

Dong and Seismic risk and Consideration of aftershocks has a Seismic, post-event
PGA
Frangopol 2015 resilience of bridges great influence on bridge functionality consequence analysis

197
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

under mainshocks and and repair cost, and hence on seismic and bridge restoration
aftershocks resilience parameters

Damage accumulation Damage index conditioned to the


Ghosh et al. in bridges due to number of shocks are proposed to
PGA __
2015 mainshock and a series predict bridge damage after future
of aftershocks earthquakes with multiple shocks

Enhanced seismic
performance of bridge Repair of a severely damaged bridge
Fakharifar et al. piers repaired with pier during mainshock helps to
SA __
2015 FRP/steel after improve its aftershock collapse
mainshock and prior to capacity only to a little extent;
aftershocks

Reliability of Bridge reliability during aftershock


Structural parameters,
Chiu and Arista mainshock-damaged decreases with (i) decreasing distance
__ source location of
2017 bridge piers under between site and source and (ii)
aftershock
aftershocks increasing number of aftershocks

Exceedance probabilities of a higher Parameters related to


Chiu et al. 2018 Aftershock-induced PGA
damage state of bridge piers increase aftershock hazard

198
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

seismic hazard for with the time gap between main and
seismic design of bridge aftershocks
piers

Multihazard risk
Mean Risks due to hurricane and earthquake
Earthquake Kameshwar and assessment based on
annual hazards are combined to obtain the __
and hurricane Padgett 2014 parameterised fragility
frequency total risk
approach

Parameters related to wave and surge


Parameter sensitivity on
are found to be most critical; Among Wave, surge and
Ataei et al. 2010 the damage of coastal __
structural parameters, deck mass is the structural parameters
bridges
most critical for bridge response

Effect of submersion The maximum uplift wave force is


Storm surge Wave
depth of a bridge due to observed to occur when the water __
and wave Xiao et al. 2010 height,
storm surge on the level due to storm surge reached the
period
maximum wave force top of bridge deck

Limit state capacities of Failure probabilities of global system Structural and


Ataei and Wave height
bridges under hurricane damage conditioned to component- geotechnical
Padgett 2013 and period,
surge and wave level demand are estimated parameters

199
Intensity Uncertain
Multihazard Reference Objectives Major Observation/Conclusion
Parameters parameters, if any

surge
elevation

Relative
Surrogate models for surge
Ataei and Fragilities are calculated for deck
generating fragilities of elevation, Structural parameters
Padgett 2015a unseating failure mode of bridges
costal bridges maximum
wave height

Relative
Effect of modelling Sensitivity of bridge vulnerability, for
surge
Ataei and parameters on deck uplift and unseating, on various Structural and fluid
elevation,
Padgett 2015b vulnerability of coastal fluid-structure integration model parameters
maximum
bridges parameters is studied
wave height

Impacts of storm surge Wave and Failure probabilities of various bridges


Anarde et al.
on coastal bridge surge are calculated for various levels of __
2018
infrastructures heights storm surge

200
APPENDIX - B
Overview of projected flood discharge for the San Joaquin river in
California

This Appendix provides the background information on the projected flood discharge
data for the period between 2012-2099 presented in the Fig. 3.7. Global climate projections are
generated by the World Climate Research Programme (WCRP) through its Coupled Model
Intercomparison Project Phase 5 (CMIP5) involving GCM simulations resulting from
modeling efforts from across the world. This state-of-the-art intercomparison project uses
updated greenhouse emission pathways depicted by representative concentration pathways
(RCP): RCP 2.6, RCP 4.5, RCP 6.0 and RCP 8.5, where the numbers indicate the net
anthropogenic radiative forcing at the top of the atmosphere at the end of the twenty-first
century. Meteorological variables such as precipitation and temperature, simulated by the
GCMs, are typically at much coarser resolution and lesser accuracy than necessary in
hydrological studies. Statistical downscaling has been commonly used to address this scale
mismatch. In this study, bias correction and spatial disaggregation (BCSD)-based statistical
downscaling (Wood et al. 2004) is used to obtain the climate projections. These daily climate
projections are used to drive the spatially distributed variable infiltration capacity (VIC)
hydrological model (Liang et al. 1994). BCSD CMIP5 climate (Bureau of Reclamation 2013)
and hydrological (Brekke et al. 2014) projections for the US are archived and available at
BCSD (2020) for 97 combinations of models and future emission scenarios. In this study,
annual maximum flows are calculated from these daily projections.

The empirical cumulative distribution functions (CDFs) of annual maximum discharge


in the San Joaquin River from these projections are evaluated with respect to the observations
for the period of overlap (1950–2011), as shown in Fig. B1 (a). Although the CMIP5 BCSD
climate has the same monthly climatology (Bureau of Reclamation 2013) as observations
(Maurer et al. 2002), it may be observed that the peak discharge shows systematic bias in the
climate projections (shown in Fig. A1 (a)). A possible reason for this could be the monthly to
annual scale of discharges that have been used to calibrate the VIC model (Brekke et al. 2014).
To circumvent this limitation, a quantile mapping–based bias correction (Hamlet et al.2010;
Mondal and Mujumdar 2016) is further adopted in this study to obtain future projections of
annual maximum flows. The CDFs of the bias corrected peak discharges show good match
with the observations (shown in Fig. A1 (b)).

201
Note that each of these projections reveal plausible future scenarios and all of them may
be equally likely (IPCC 2007). However, it is infeasible to account for all of them in the future
design or adaptation strategies. Therefore, multimodel combinations such as the mean, median,
or maximum of the projections are considered inearlier studies (e.g.,Mondal and Mujumdar
2016). In this study, the maximum value of peak flow out of all the projections in each year
yields unrealistically large flood magnitudes. Also, each of these projections are derived from
model simulations that conserve physics and combining them into a mean or a median may not
yield physically consistent results from year to year. Therefore, a representative projection,
namely the French Global ClimateModel (CNRM) simulations, corresponding to the RCP 8.5
is chosen for further analysis. This projection shows a reasonably good match with
observations and also visually lies close to the median of the range of projections during the
historical period of overlap (Fig. B.1 (b)).

Figure B.1 Cumulative distribution function (CDF) plot for annual peak discharge of the river
from observations (black) and (a) raw and (b) bias-corrected projections (other colours)
during the period of overlap between observations and model simulations (1950-2011)

202
APPENDIX - C
Table C.1 Ground motions used to match seismic hazard spectrum in Nepal compatible with 10% exceedance in 50 years.

Ground motion Moment PGA (g)


PEER ID Earthquake name Year Recording station
ID magnitude (Scaled)
Castaic - Old Ridge
1 RSN 41 Lytle Creek 1970 5.3 0.23
Route
Cedar Springs
2 RSN 42 Lytle Creek 1970 5.3 0.27
Pumphouse
Puddingstone Dam
3 RSN 48 Lytle Creek 1970 5.3 0.26
(Abutment)
Wrightwood - 6074
4 RSN 50 Lytle Creek 1970 5.3 0.28
Park Dr
Cedar Sprin - Allen
5 RSN 59 San Fernando 1971 6.6 0.23
Ranch
Gormon - Oso
6 RSN 65 San Fernando 1971 6.6 0.27
Pump Plant
7 RSN 66 San Fernando 1971 6.6 Hemet Fire Station 0.22
LA - Hollywood
8 RSN 68 San Fernando 1971 6.6 0.23
Stor FF
9 RSN 72 San Fernando 1971 6.6 Lake Hughes #4 0.31
Palmdale Fire
10 RSN 78 San Fernando 1971 6.6 0.23
Station
Pasadena - Old
11 RSN 80 San Fernando 1971 6.6 0.24
Seismo Lab

203
Ground motion Moment PGA (g)
PEER ID Earthquake name Year Recording station
ID magnitude (Scaled)
Pasadena - Old
11 RSN 80 San Fernando 1971 6.6 0.24
Seismo Lab
Wheeler Ridge -
12 RSN 92 San Fernando 1971 6.6 0.30
Ground
Whittier Narrows
13 RSN 93 San Fernando 1971 6.6 0.23
Dam
Wrightwood - 6074
14 RSN 94 San Fernando 1971 6.6 0.23
Park Dr
15 RSN 126 Gazli 1976 6.8 Karakyr 0.31
16 RSN 127 Fruili 1976 5.5 Buia 0.23
17 RSN 139 Tabas 1978 7.3 Dayhook 0.23
18 RSN 140 Tabas 1978 7.3 Ferdows 0.24
19 RSN 308 Taiwan 1981 5.9 Smart1 I12 0.22
20 RSN 309 Taiwan 1981 5.9 Smart1 M01 0.21

204
Table C.2 Ground motions used to match seismic hazard spectrum in Nepal compatible with 5% exceedance in 50 years.

Ground motion Moment PGA (g)


PEER ID Earthquake name Year Recording station
ID magnitude (Scaled)
Castaic - Old Ridge
21 RSN 41 Lytle Creek 1970 5.3 0.34
Route
Puddingstone Dam
22 RSN 48 Lytle Creek 1970 5.3 0.38
(Abutment)
Wrightwood - 6074
23 RSN 50 Lytle Creek 1970 5.3 0.40
Park Dr
Castaic - Old Ridge
24 RSN 57 San Fernando 1971 6.6 0.34
Route
Cedar Springs -
25 RSN 59 San Fernando 1971 6.6 0.33
Allen Ranch
Gormon - Oso
26 RSN 65 San Fernando 1971 6.6 0.38
Pump Plant
27 RSN 66 San Fernando 1971 6.6 Hemet Fire Station 0.32
LA - Hollywood
28 RSN 68 San Fernando 1971 6.6 0.33
Stor FF
29 RSN 72 San Fernando 1971 6.6 Lake Hughes #4 0.45
Palmdale Fire
30 RSN 78 San Fernando 1971 6.6 0.33
Station
Pasadena - Old
31 RSN 80 San Fernando 1971 6.6 0.34
Seismo Lab
Whittier Narrows
32 RSN 93 San Fernando 1971 6.6 0.33
Dam

205
Ground motion Moment PGA (g)
PEER ID Earthquake name Year Recording station
ID magnitude (Scaled)
Wrightwood - 6074
33 RSN 94 San Fernando 1971 6.6 0.32
Park Dr
34 RSN 126 Gazli 1976 6.8 Karakyr 0.44
35 RSN 127 Fruili 1976 5.5 Buia 0.32
36 RSN 139 Tabas 1978 7.3 Dayhook 0.33
37 RSN 140 Tabas 1978 7.3 Ferdows 0.35
38 RSN 143 Tabas 1978 7.3 Tabas 0.27
39 RSN 308 Taiwan 1981 5.9 Smart1 I12 0.32
40 RSN 309 Taiwan 1981 5.9 Smart1 M01 0.31

Table C.3 Ground motions used to match seismic hazard spectrum in Nepal compatible with 2% exceedance in 50 years.

Ground motion Earthquake Moment PGA (g)


PEER ID Year Recording station
ID name magnitude (Scaled)
Castaic - Old Ridge
41 RSN 41 Lytle Creek 1970 5.3 0.51
Route
Puddingstone Dam
42 RSN 48 Lytle Creek 1970 5.3 0.57
(Abutment)
Wrightwood - 6074
43 RSN 50 Lytle Creek 1970 5.3 0.60
Park Dr
Castaic - Old Ridge
44 RSN 57 San Fernando 1971 6.6 0.51
Route

206
Ground motion Earthquake Moment PGA (g)
PEER ID Year Recording station
ID name magnitude (Scaled)
Cedar Springs -
45 RSN 59 San Fernando 1971 6.6 0.50
Allen Ranch
Gormon - Oso
46 RSN 65 San Fernando 1971 6.6 0.57
Pump Plant
47 RSN 66 San Fernando 1971 6.6 Hemet Fire Station 0.48
LA - Hollywood
48 RSN 68 San Fernando 1971 6.6 0.49
Stor FF
49 RSN 72 San Fernando 1971 6.6 Lake Hughes #4 0.68
Palmdale Fire
50 RSN 78 San Fernando 1971 6.6 0.49
Station
Pasadena - Old
51 RSN 80 San Fernando 1971 6.6 0.52
Seismo Lab
Whittier Narrows
52 RSN 93 San Fernando 1971 6.6 0.50
Dam
Wrightwood - 6074
53 RSN 94 San Fernando 1971 6.6 0.49
Park Dr
54 RSN 126 Gazli 1976 6.8 Karakyr 0.67
55 RSN 127 Fruili 1976 5.5 Buia 0.49
56 RSN 139 Tabas 1978 7.3 Dayhook 0.50
57 RSN 140 Tabas 1978 7.3 Ferdows 0.53
58 RSN 307 Taiwan 1981 5.9 Smart1 I06 0.42
59 RSN 308 Taiwan 1981 5.9 Smart1 I12 0.48

207
Ground motion Earthquake Moment PGA (g)
PEER ID Year Recording station
ID name magnitude (Scaled)
60 RSN 309 Taiwan 1981 5.9 Smart1 M01 0.46

Table C.4 Recorded ground motions in Nepal obtained from the Center for Engineering Strong Motion Data

Ground motion Earthquake Moment Recording PGA (g)


PEER ID Year
ID name magnitude station (Scaled)
Municipality
61 KTP Gorkha 2015 7.8 0.26
office, Kirtipur
Tirubhuvan
62 TVU Gorkha 2015 7.8 University, 0.23
Kirtipur
Kanti Path,
63 KATNP Gorkha 2015 7.8 0.16
Kathmandu
Pulchowk
64 PTN Gorkha 2015 7.8 0.15
Campus, Patan
Municipality
61 KTP Gorkha 2015 7.8 0.26
office, Kirtipur

208
Table C.5 Selected ground motions for seismic fragility analysis of RC highway bridge in the state of Gujarat, India

Ground Earthquake Moment PGA (g)


PEER ID Year Recording station
motion ID. name magnitude (Scaled)
San
1 68 1971 6.6 LA - Hollywood Stor FF 0.10
Fernando
San
2 70 1971 6.6 Lake Hughes #1 0.14
Fernando
San
3 88 1971 6.6 Santa Felita Dam (Outlet) 0.15
Fernando
4 138 Tabas (Iran) 1978 7.3 Boshrooyeh 0.10
5 322 Coalinga 1983 6.4 Cantua Creek School 0.11
6 359 Coalinga 1983 6.4 Parkfield - Vineyard Cany 1E 0.15
7 963 Northridge 1994 6.7 Castaic - Old Ridge Route 0.14
8 974 Northridge 1994 6.7 Glendale - Las Palmas 0.30
9 985 Northridge 1994 6.7 LA - Baldwin Hills 0.13
10 987 Northridge 1994 6.7 LA - Centinela St 0.17
11 998 Northridge 1994 6.7 LA - N Westmoreland 0.21
12 1003 Northridge 1994 6.7 LA - Saturn St 0.18
13 1057 Northridge 1994 6.7 Playa Del Rey - Saran 0.10
14 2495 Chi-Chi 1999 6.2 CHY080 0.16

209
Ground Earthquake Moment PGA (g)
PEER ID Year Recording station
motion ID. name magnitude (Scaled)
15 2507 Chi-Chi 1999 6.2 CHY101 0.10
16 2623 Chi-Chi 1999 6.2 TCU072 0.19
17 2650 Chi-Chi 1999 6.2 TCU116 0.10
18 2654 Chi-Chi 1999 6.2 TCU120 0.10
19 3467 Chi-Chi 1999 6.3 TCU065 0.12
20 3468 Chi-Chi 1999 6.3 TCU067 0.10
21 3505 Chi-Chi 1999 6.3 TCU125 0.14
22 4226 Niigata 2004 6.6 NIGH09 0.22
23 4849 Chuetsu 2007 6.8 Kubikiku Hyakken Joetsu City 0.11
24 4855 Chuetsu 2007 6.8 Sanjo 0.10
Hinodecho Yoshida Tsubame
25 4880 Chuetsu 2007 6.8 City
0.10

26 5267 Chuetsu 2007 6.8 NIG021 0.16


28 5284 Chuetsu 2007 6.8 NIGH11 0.12
29 5620 Iwate 2008 6.9 IWT012 0.20
30 5663 Iwate 2008 6.9 MYG004 0.17

210
Table C.6 SRSS of vertical spectral acceleration of selected ground motions at different life-cycle year of bridge

PEER SRSS of vertical spectral acceleration at first and second vertical vibration mode of
Record bridge (g)
S. No
Sequence
Number Pristine 25 years 50 years 75 years 100 years

1 68 0.36 0.36 0.36 0.28 0.30


2 70 0.44 0.44 0.44 0.39 0.39
3 88 0.45 0.45 0.45 0.45 0.42
4 138 0.24 0.24 0.24 0.26 0.24
5 322 0.16 0.16 0.16 0.14 0.13
6 359 0.56 0.56 0.56 0.59 0.55
7 963 0.16 0.16 0.16 0.17 0.17
8 974 0.31 0.31 0.31 0.27 0.25
9 985 0.24 0.24 0.24 0.26 0.28
10 987 0.29 0.29 0.29 0.22 0.22
11 998 0.31 0.31 0.31 0.30 0.30
12 1003 0.31 0.31 0.31 0.41 0.41
13 1057 0.34 0.34 0.34 0.27 0.31
14 2495 0.84 0.84 0.84 0.43 0.42
15 2507 0.22 0.22 0.22 0.18 0.17

211
PEER SRSS of vertical spectral acceleration at first and second vertical vibration mode of
Record bridge (g)
S. No
Sequence
Number Pristine 25 years 50 years 75 years 100 years
16 2623 0.34 0.34 0.34 0.23 0.21
17 2650 0.31 0.31 0.31 0.39 0.34
18 2654 0.33 0.33 0.33 0.21 0.21
19 3467 0.26 0.26 0.26 0.31 0.28
20 3468 0.24 0.24 0.24 0.23 0.23
21 3505 0.56 0.56 0.56 0.44 0.48
22 4226 0.21 0.21 0.21 0.24 0.23
23 4849 0.26 0.26 0.26 0.20 0.16
24 4855 0.81 0.81 0.81 0.42 0.40
25 4880 0.21 0.21 0.21 0.29 0.32
26 5267 0.44 0.44 0.44 0.62 0.72
27 5270 1.80 1.80 1.80 1.10 1.14
28 5284 0.41 0.41 0.41 0.49 0.48
29 5620 0.42 0.42 0.42 0.30 0.29
30 5663 0.45 0.45 0.45 0.27 0.29

212
PUBLICATIONS

Referred International Journals

1) Banerjee S., Sharanbaswa B. Vishwanath, Devendiran D.K., (2019). “Multihazard


resilience of highway bridges and bridge networks: a review.” Structure and Infrastructure
Engineering, Taylor and Francis, 15(12), 1694-1714.
2) Devendiran, D. K., Banerjee, S., and Mondal, A. (2021). “Impact of climate change on
multihazard performance of river-crossing bridges: Risk, resilience, and
adaptation.” Journal of Performance of Constructed Facilities, 35(1), 04020127.
3) Devendiran, D. K., and Banerjee, S. (2023). “Influence of Combined Corrosion–Fatigue
Deterioration on Life-Cycle Resilience of RC Bridges.” Journal of Bridge
Engineering, 28(5), 04023014.
4) Devendiran, D. K., and Banerjee, S. (2023). “Effect of bridge design parameters on multi-
hazard performance of river-crossing bridges.” Proceedings of the Institution of Civil
Engineers-Structures and Buildings, 1-19.

Manuscript under review/preparation

5) Devendiran, D. K., and Banerjee, S. “Contribution of vertical ground motion on seismic


response of reinforced concrete bridges in the presence of corrosion-fatigue degradation”
[Status: Submitted]
6) Devendiran, D. K., and Banerjee, S. “Life-cycle based seismic vulnerability assessment of
bridges in the presence of corrosion-fatigue degradation” [Status: In-preparation]

Referred International Conference Proceedings

7) Devendiran, D. K., and Banerjee, S. (2021). “Impact of climate change on multihazard


vulnerability of highway bridges.” In Bridge Maintenance, Safety, Management, Life-
Cycle Sustainability and Innovations (pp. 3814-3819). CRC Press.

213
REFERENCES

AASHTO (1990). Guide specifications for fatigue evaluation of existing steel bridges. Washington, DC:
AASHTO

AASHTO (2008). LRFD bridge design specifications. Washington, DC: AASHTO.

AASHTO (American Association of State Highways and Transportation Officials) (2017).


LRFD Bridge Design Specifications. AASHTO, Washington, DC, USA.

ACI (2005). Building code requirement for structural concrete (ACI 318-05) and commentary
(ACI 318R-05). American Concrete Institute.
Adhikari, R., Gautam, D., Jha, P., Aryal, B., Ghalan, K., Rupakhety, R., ... and Motra, G.
(2019). “Bridging multi-hazard vulnerability and sustainability: approaches and
applications to nepali highway bridges.” Resilient Structures and Infrastructure, 361-
378.

Agostinacchio, M., Ciampa, D., and Olita, S. (2014). “The vibrations induced by surface
irregularities in road pavements – A Matlab approach.” European Transport Research
Review, 6(3), 267-275.

Ahn, W., and Reddy, D. V. (2001). “Galvanostatic testing for the durability of marine concrete
under fatigue loading.” Cement and Concrete Research, 31(3), 343-349.

Akiyama, M., and Frangopol, D. M. (2014). “Long-term seismic performance of RC structures


in an aggressive environment: Emphasis on bridge piers.” Structure and Infrastructure
Engineering, 10(7), 865-879.
Akiyama, M., Frangopol, D. M., and Matsuzaki, H. (2011). “Life‐cycle reliability of RC bridge
piers under seismic and airborne chloride hazards.” Earthquake Engineering &
Structural Dynamics, 40(15), 1671-1687.
Akiyama, M., Frangopol, D. M., Arai, M., and Koshimura, S. (2013). “Reliability of bridges
under tsunami hazards: Emphasis on the 2011 Tohoku-oki earthquake.” Earthquake
Spectra, 29 (1_suppl), 295-314.

Al-Attraqchi, A. Y., Hashemi, M. J., and Al-Mahaidi, R. (2022). “Loss assessment of rigid-
frame bridges under horizontal and vertical ground motions.” In Structures (Vol. 35,
pp. 243-259). Elsevier.

214
Alessandri, S., Giannini, R., and Paolacci, F. (2013). “Aftershock risk assessment and the
decision to open traffic on bridges.” Earthquake Engineering & Structural
Dynamics, 42(15), 2255-2275.
Alipour, A., and Shafei, B. (2012). “Performance assessment of highway bridges under
earthquake and scour effects.” In Proceedings of the 15th world conference on
earthquake engineering (pp. 24-28).

Alipour, A., and Shafei, B. (2016). “Seismic resilience of transportation networks with
deteriorating components.” Journal of Structural Engineering, 142(8), C4015015.
Alipour, A., Shafei, B., and Shinozuka, M. (2013). “Reliability-based calibration of load and
resistance factors for design of RC bridges under multiple extreme events: Scour and
earthquake.” Journal of Bridge Engineering, 18(5), 362-371.

Ambraseys, N. N., and Douglas, J. (2003). “Near-field horizontal and vertical earthquake
ground motions.” Soil dynamics and earthquake engineering, 23(1), 1-18.
Ambraseys, N. N., Simpson, K. U., and Bommer, J. J. (1996). “Prediction of horizontal
response spectra in Europe.” Earthquake Engineering & Structural Dynamics, 25(4),
371-400.
Anarde, K. A., Kameshwar, S., Irza, J. N., Nittrouer, J. A., Lorenzo-Trueba, J., Padgett, J. E.,
... and Bedient, P. B. (2018). “Impacts of hurricane storm surge on infrastructure
vulnerability for an evolving coastal landscape.” Natural Hazards Review, 19(1),
04017020.
Anderson, C., Claman, D., and Mantilla, R. (2015). Iowa's Bridge and Highway Climate
Change and Extreme Weather Vulnerability Assessment Pilot (No. HEPN-707). Iowa
State University. Institute for Transportation.

Andrić, J. M., and Lu, D. G. (2017). “Fuzzy methods for prediction of seismic resilience of
bridges.” International Journal of Disaster Risk Reduction, 22, 458-468.

Antolik, M., and Dreger, D. S. (2003). “Rupture process of the 26 January 2001 M w 7.6 Bhuj,
India, earthquake from teleseismic broadband data.” Bulletin of the Seismological
Society of America, 93(3), 1235-1248.
API (2003). Recommended practice for planning, designing and constructing fixed offshore
platforms: Working stress design. Washington, DC: API.

215
Aryan, H., and Ghassemieh, M. (2020). “Numerical assessment of vertical ground motion
effects on highway bridges.” Canadian Journal of Civil Engineering, 47(7), 790-800.
Ataei, N., & Padgett, J. E. (2015a). Fragility surrogate models for coastal bridges in hurricane
prone zones. Engineering Structures, 103, 203-213.
Ataei, N., & Padgett, J. E. (2015b). Influential fluid–structure interaction modelling parameters
on the response of bridges vulnerable to coastal storms. Structure and Infrastructure
Engineering, 11(3), 321-333.
Ataei, N., and Padgett, J. E. (2013). “Limit state capacities for global performance assessment
of bridges exposed to hurricane surge and wave.” Structural safety, 41, 73-81.

ATC (1985). Earthquake damage evaluation data for California. Rep. ATC-13. Redwood
City, CA: ATC.

Aviram, A., Mackie, K. R., and Stojadinovic, B. (2008). “Effect of abutment modeling on the
seismic response of bridge structures.” Earthquake Engineering and Engineering
Vibration, 7(4), 395.

Aygün, B., Dueñas-Osorio, L., Padgett, J. E., and DesRoches, R. (2011). “Efficient longitudinal
seismic fragility assessment of a multispan continuous steel bridge on liquefiable
soils.” Journal of Bridge Engineering, 16(1), 93-107.
Ayyub, B. M., and Popescu, C. (2003). “Risk-based expenditure allocation for infrastructure
improvement.” Journal of Bridge Engineering, 8(6), 394-404.

Banerjee, S., and Ganesh Prasad, G. (2013). “Seismic risk assessment of reinforced concrete
bridges in flood-prone regions.” Structure and Infrastructure Engineering, 9(9), 952-
968.

Banerjee, S., and Shinozuka, M. (2007). “Nonlinear static procedure for seismic vulnerability
assessment of bridges.” Comp er‐Aided Civi nd Infr s r c re Engineering, 22(4),
293-305.

Banerjee, S., Vishwanath, B. S., and Devendiran, D. K. (2019). “Multihazard resilience of


highway bridges and bridge networks: A review.” Structure and Infrastructure
Engineering, 15(12), 1694-1714.

Barnett, T. P., Pierce, D. W., Hidalgo, H. G., Bonfils, C., Santer, B. D., Das, T., ... and
Dettinger, M. D. (2008). “Human-induced changes in the hydrology of the western
United States.” science, 319(5866), 1080-1083.

216
Basöz, N. I., and Kiremidjian, A. S. (1998). “Evaluation of bridge damage data from the Loma
Prieta and Northridge, California earthquakes.” In Evaluation of bridge damage data
from the Loma Prieta and Northridge, California earthquakes (pp. 167-167).

Basquin, O. H. (1910). “The exponential law of endurance tests.” In Proc Am Soc Test
Mater (Vol. 10, pp. 625-630).

Bastidas-Arteaga, E. (2009). Probabilistic service life modeling of RC structures subjected to


the combined effect of chloride-induced corrosion and cyclic loading (Doctoral
dissertation, Universidad de los Andes).

Bastidas-Arteaga, E., Bressolette, P., Chateauneuf, A., and Sánchez-Silva, M. (2009).


“Probabilistic lifetime assessment of RC structures under coupled corrosion–fatigue
deterioration processes.” Structural safety, 31(1), 84-96.
BCSD (Bias Correction and Spatial Disaggregation) (2020). Downscaled CMIP3 and CMIP5
climate and hydrology projections. Accessed May 31, 2020. http://gdo-
dcp.ucllnl.org/pub/dcp/archive/cmip5/hydro.
Berry, M. P., and Eberhard, M. O. (2004). Performance models for flexural damage in
reinforced concrete columns. Pacific Earthquake Engineering Research Center, Univ.
of California, Berkeley.

Berry, M. P., and Eberhard, M. O. (2005). “Practical performance model for bar
buckling.” Journal of Structural Engineering, 131(7), 1060-1070.

Bertolini, L., Elsener, B., Pedeferri, P., Redaelli, E., and Polder, R. B. (2013). Corrosion of
steel in concrete: prevention, diagnosis, repair. John Wiley & Sons.
Bilham, R. (2019). “Himalayan earthquakes: a review of historical seismicity and early 21st
century slip potential.” Geological Society, London, Special Publications, 483(1), 423-
482.

Biondini, F., Camnasio, E., and Palermo, A. (2014). “Lifetime seismic performance of concrete
bridges exposed to corrosion.” Structure and Infrastructure Engineering, 10(7), 880-
900.
Biondini, F., Camnasio, E., and Titi, A. (2015). “Seismic resilience of concrete structures under
corrosion.” Earthquake Engineering & Structural Dynamics, 44(14), 2445-2466.

217
BIS (2002). IS 1893 (part 1): 2002—Indian standard criteria for earthquake resistant design
of structures, part 1: general provisions and buildings (Fifth Revision). Bureau of
Indian Standards, New Delhi.

Biswas, S. K. (1987). “Regional tectonic framework, structure and evolution of the western
marginal basins of India.” Tectonophysics, 135(4), 307-327.

Blikharskyy, Y., Selejdak, J., and Kopiika, N. (2021). “Corrosion fatigue damages of rebars
under loading in time.” Materials, 14(12), 3416.

BMS (Bridge Management System) (2022) Bridge Inventory Report. See


http://bms.softavi.com/dashboard/guest_report_bi (accessed28/08/2022).

Bocchini, P., and Frangopol, D. M. (2012a). “Restoration of bridge networks after an


earthquake: Multicriteria intervention optimization.” Earthquake spectra, 28(2), 427-
455.

Bocchini, P., and Frangopol, D. M. (2012b). “Optimal resilience-and cost-based post disaster
intervention prioritization for bridges along a highway segment.” Journal of Bridge
Engineering, 17(1), 117-129.

Bocchini, P., Frangopol, D. M., Ummenhofer, T., and Zinke, T. (2014). “Resilience and
sustainability of civil infrastructure: Toward a unified approach.” Journal of
Infrastructure Systems, 20(2), 04014004.

Bouwer, L. M. (2019). “Observed and projected impacts from extreme weather events:
implications for loss and damage.” Loss and damage from climate change: concepts,
methods and policy options, 63-82.

Brekke, L., A. Wood, and T. Pruitt (2014). Downscaled CMIP3 and CMIP5 hydrology
projections: Release of hydrology projections, comparison with preceding information,
and summary of user needs. Boulder, CO: National Center for Atmospheric Research.
Bruneau, M., Chang, S. E., Eguchi, R. T., Lee, G. C., O'Rourke, T. D., Reinhorn, A. M., ... &
Von Winterfeldt, D. (2003). “A framework to quantitatively assess and enhance the
seismic resilience of communities.” Earthquake spectra, 19(4), 733-752.

Bureau of Reclamation (2013). Downscaled CMIP3 and CMIP5 climate projections: Release
of downscaled CMIP5 climate projections, com-parison with preceding information,
and summary of user needs. Denver: US Dept. of the Interior, Bureau of Reclamation,
Technical Service Center.

218
Button, M. R., Cronin, C. J., and Mayes, R. L. (2002). Effect of vertical motions on seismic
response of highway bridges. Journal of structural engineering, 128(12), 1551-1564.

Byers, W. G., Marley, M. J., Mohammadi, J., Nielsen, R. J., and Sarkani, S. (1997). “Fatigue
reliability reassessment applications: State-of-the-art paper.” Journal of Structural
Engineering, 123(3), 277-285.

Caltrans (2018). Caltrans climate change vulnerability assessments. District 4 Technical


report, Caltrans, California.

Caltrans (2021). Corrosion guidelines. Version 3.2. Technical report, Sacramento, CA: Caltrans.

Cayan, D. R., Maurer, E. P., Dettinger, M. D., Tyree, M., and Hayhoe, K. (2008). “Climate
change scenarios for the California region.” Climatic change, 87, 21-42.

CEN (European Committee for Standardization) (2005). Design of steel structure—Part 1–9: Fatigue.
Eurocode 3. EN1993-1-9. Brussels, Belgium: CEN.

Chandrasekaran, S., and Banerjee, S. (2016). “Retrofit optimization for resilience enhancement
of bridges under multihazard scenario.” Journal of Structural Engineering, 142(8),
C4015012.

Chang, K. C., Sung, Y. C., Liu, K. Y., Wang, P. H., Lee, Z. K., Lee, L. S., and Witarto. (2014).
“Seismic performance of an existing bridge with scoured caisson
foundation.” Earthquake Engineering and Engineering Vibration, 13, 151-165.

Chang, S. E., Shinozuka, M., and Moore, J. E. (2000). “Probabilistic earthquake scenarios:
extending risk analysis methodologies to spatially distributed systems.” Earthquake
Spectra, 16(3), 557-572.

Chaulagain, H., Rodrigues, H., Varum, H., Silva, V., & Gautam, D. (2017). “Generation of
spectrum-compatible acceleration time history for Nepal.” Comptes Rendus
Geoscience, 349(5), 198-201.

Chiu, C. K., and Arista, C. P. (2017). “Serviceability-related reliability for mainshock-damaged


reinforced concrete piers considering the aftershock-induced seismic hazards.” Natural
Hazards, 87, 1333-1359.
Chiu, C. K., Liao, I. H., and Jean, W. Y. (2018). “Seismic design requirements for reinforced
concrete piers considering aftershock-induced seismic hazard.” Structure and
Infrastructure Engineering, 14(9), 1244-1256.

219
Choe, D. E., Gardoni, P., and Rosowsky, D. (2010). “Fragility increment functions for
deteriorating reinforced concrete bridge columns.” Journal of engineering
mechanics, 136(8), 969-978.

Choe, D. E., Gardoni, P., Rosowsky, D., and Haukaas, T. (2009). “Seismic fragility estimates
for reinforced concrete bridges subject to corrosion.” Structural Safety, 31(4), 275-283.
Choi, E. (2002). Seismic analysis and retrofit of mid-America bridges. PhD Dissertation,
Georgia Institute of Technology.

Choine, M. N., O’Connor, A., and Padgett, J. E. (2013). “A seismic reliability assessment of
reinforced concrete integral bridges subject to corrosion.” In Key Engineering
Materials (Vol. 569, pp. 366-373). Trans Tech Publications Ltd.
Chopra, S., Kumar, D., Rastogi, B. K., Choudhury, P., and Yadav, R. B. S. (2012).
“Deterministic seismic scenario for Gujarat region, India.” Natural hazards, 60, 517-
540.

Chotickai, P., and Bowman, M. D. (2006). “Truck models for improved fatigue life predictions
of steel bridges.” Journal of Bridge Engineering, 11(1), 71-80.

Choudhury, P., Chopra, S., and Roy, K. S. (2016). “Site classification for strong motion stations
in Gujarat, India using response spectral ratio.” Soil Dynamics and Earthquake
Engineering, 87, 138-150.
Cimellaro, G. P., Renschler, C., Reinhorn, A. M., and Arendt, L. (2016). “PEOPLES: a
framework for evaluating resilience.” Journal of Structural Engineering, 142(10),
04016063.

Coffin Jr, L. F. (1954). “A study of the effects of cyclic thermal stresses on a ductile
metal.” Transactions of the American Society of Mechanical engineers, 76(6), 931-949.

Coin News (2019). US Inflation Calculator. Accessed May 31, 2019.


https://www.usinflationcalculator.com/inflation/current-inflation-rates/

Collier, C. J., and Elnashai, A. S. (2001). “A procedure for combining vertical and horizontal
seismic action effects.” Journal of Earthquake Engineering, 5(04), 521-539.

Cook, W., Barr, P. J., and Halling, M. W. (2015). “Bridge failure rate.” Journal of Performance
of Constructed Facilities, 29(3), 04014080.

220
Costa, A., and Appleton, J. (1999). “Chloride penetration into concrete in marine
environment—Part I: Main parameters affecting chloride penetration.” Materials and
Structures, 32, 252-259.
CSA (2014). Canadian Highway Bridge Design Code. CSA International, Ontario, Canada.

CSA (Canadian Standards Association) (2014). CSA S6-19: Canadian highway bridge design
code. CSA, Mississauga, Canada.

Das, T., Maurer, E. P., Pierce, D. W., Dettinger, M. D., and Cayan, D. R. (2013). “Increases in
flood magnitudes in California under warming climates.” Journal of Hydrology, 501,
101-110.

Decò, A., and Frangopol, D. M. (2011). “Risk assessment of highway bridges under multiple
hazards.” Journal of Risk Research, 14(9), 1057-1089.

Decò, A., and Frangopol, D. M. (2013). “Life-cycle risk assessment of spatially distributed
aging bridges under seismic and traffic hazards.” Earthquake Spectra, 29(1), 127-153.

Decò, A., Bocchini, P., and Frangopol, D. M. (2013). “A probabilistic approach for the
prediction of seismic resilience of bridges.” Earthquake Engineering & Structural
Dynamics, 42(10), 1469-1487.

Dehghanpoor, A., Thambiratnam, D., Zhang, W., Chan, T., and Taciroglu, E. (2021). “An
extended probabilistic demand model with optimal intensity measures for seismic
performance characterization of isolated bridges under coupled horizontal and vertical
motions.” Bulletin of Earthquake Engineering, 19, 2291-2323.
Deng, L., and Cai, C. S. (2009). “Identification of parameters of vehicles moving on
bridges.” Engineering Structures, 31(10), 2474-2485.

Deng, P., Zhang, C., Pei, S., and Jin, Z. (2018). “Modeling the impact of corrosion on seismic
performance of multi-span simply-supported bridges.” Construction and Building
Materials, 185, 193-205.

Devendiran, D. K., and Banerjee, S. (2023). “Influence of Combined Corrosion–Fatigue


Deterioration on Life-Cycle Resilience of RC Bridges.” Journal of Bridge
Engineering, 28(5), 04023014.

Devendiran, D. K., Banerjee, S., and Mondal, A. (2021). “Impact of climate change on
multihazard performance of river-crossing bridges: Risk, resilience, and
adaptation.” Journal of Performance of Constructed Facilities, 35(1), 04020127.

221
Devkota, L. P., and Gyawali, D. R. (2015). “Impacts of climate change on hydrological regime
and water resources management of the Koshi River Basin, Nepal.” Journal of
Hydrology: Regional Studies, 4, 502-515.

Dey, A., and Sil, A. (2021). “Advanced corrosion-rate model for comprehensive seismic
fragility assessment of chloride affected RC bridges located in the coastal region of
India.” In Structures (Vol. 34, pp. 947-963). Elsevier.

Dikanski, H., Hagen-Zanker, A., Imam, B., and Avery, K. (2016, June). “Climate change
impacts on railway structures: bridge scour.” In Proceedings of the Institution of Civil
Engineers-Engineering Sustainability (Vol. 170, No. 5, pp. 237-248). Thomas Telford
Ltd.

Domaneschi, M., and Martinelli, L. (2016). “Earthquake-resilience-based control solutions for


the extended benchmark cable-stayed bridge.” Journal of Structural
Engineering, 142(8), C4015009.

Dong, Y., and Frangopol, D. M. (2015). “Risk and resilience assessment of bridges under
mainshock and aftershocks incorporating uncertainties.” Engineering Structures, 83,
198-208.
Dong, Y., and Frangopol, D. M. (2016). “Probabilistic time-dependent multihazard life-cycle
assessment and resilience of bridges considering climate change.” Journal of
Performance of Constructed Facilities, 30(5), 04016034.

Dong, Y., Frangopol, D. M., and Saydam, D. (2013). “Time‐variant sustainability assessment
of seismically vulnerable bridges subjected to multiple hazards.” Earthquake
Engineering & Structural Dynamics, 42(10), 1451-1467.
DuraCrete (2000). General guidelines for durability design and redesign. Report No. BE95-
1347/R17. Prepared for DuraCrete: Probabilistic performance-based durability
design of concrete structures, The European Union – Brite EuRam III.
Dutta, A., and Mander, J. B. (1998). Capacity design and fatigue analysis of confined concrete
columns (No. MCEER-98-0007). State University of New York at Buffalo. Department
of Civil, Structural, and Environmental Engineering.

Elgamal, A., Lai, T., Yang, Z., He, L. Dynamic soil properties, seismic downhole arrays and
applications in practice. In Proceedings of the International Conference on Recent

222
Advances in Geotechnical Earthquake Engineering and Soil Dynamics, Rolla, MO,
USA, 26–31 March 2001; Volume 6.

Ellingwood, B. R. (2005). “Risk-informed condition assessment of civil infrastructure: state of


practice and research issues. “Structure and infrastructure engineering, 1(1), 7-18.

Elwood, K. J., and Eberhard, M. O. (2009). “Effective Stiffness of Reinforced Concrete


Columns.” ACI Structural Journal, 106(4).
Eslamnia, H., Malekzadeh, H., Jalali, S. A., and Moghadam, A. S. (2023). “Seismic energy
demands and optimal intensity measures for continuous concrete box-girder
bridges.” Soil Dynamics and Earthquake Engineering, 165, 107657.

Eurocode (2005). Eurocode 8: Design provisions of structures for earthquake resistance—Part


2: Bridges (prEN1998-2, Final Draft). Brussels, Belgium: CEN.

Eurocode 2 (2004). Design of concrete structures, Part 1. CEN (Comite Europeen de


Normalisation), European standard ENV 1992-1-1, Brussels; 2004.
Fakharifar, M., Chen, G., Sneed, L., and Dalvand, A. (2015). “Seismic performance of post-
mainshock FRP/steel repaired RC bridge columns subjected to
aftershocks.” Composites Part B: Engineering, 72, 183-198.
FEMA (2003). Multi-hazard loss estimation methodology, earthquake model: HAZUS-MH
MR4 technical manual. Washington, DC: FEMA.

FEMA (2013). HAZUS MR4 technical manual, Dept. of Homeland Security, Emergency Preparedness and
Response Directorate, FEMA Mitigation Division, Washington, DC.

FEMA (Federal Emergency Management Agency) (1999). HAZUS 99 Earthquake Loss


Es im ion Me hodo ogy: User’s M n . FEMA, Washington, DC, USA.

FEMA 451B (2007). NEHRP Recommended Provisions for New Buildings and Other
Structures: Training and Instructional Materials. Washington, DC, USA.

FHWA (2019). Bridge condition by highway system. Accessed October 26, 2019.
https://www.fhwa.dot.gov/bridge/britab.cfm

Franchin, P., and Pinto, P. E. (2009). “Allowing traffic over mainshock-damaged


bridges.” Journal of Earthquake Engineering, 13(5), 585-599.

223
Ganesh Prasad, G., and Banerjee, S. (2013). “The impact of flood-induced scour on seismic
fragility characteristics of bridges.” Journal of Earthquake Engineering, 17(6), 803-
828.

Ganesh Prasad, G., and Banerjee, S. (2013). “The impact of flood-induced scour on seismic
fragility characteristics of bridges.” Journal of Earthquake Engineering, 17(6), 803-
828.
Gardoni, P., and Rosowsky, D. (2011). “Seismic fragility increment functions for deteriorating
reinforced concrete bridges.” Structure and Infrastructure Engineering, 7(11), 869-
879.
Gautam, D., Rupakhety, R., and Adhikari, R. (2019). “Empirical fragility functions for Nepali
highway bridges affected by the 2015 Gorkha Earthquake.” Soil Dynamics and
Earthquake Engineering, 126, 105778.

Ge, H., Kang, L., and Tsumura, Y. (2013). “Extremely low-cycle fatigue tests of thick-walled
steel bridge piers.” Journal of Bridge Engineering, 18(9), 858-870.

Gehl, P., and D’ayala, D. (2016). “Development of Bayesian Networks for the multi-hazard
fragility assessment of bridge systems.” Structural Safety, 60, 37-46.

Ghosh, J. (2013). Parameterized seismic reliability assessment and life-cycle analysis of aging
highway bridges, Doctoral dissertation, Rice University.

Ghosh, J., and Padgett, J. E. (2010). “Aging considerations in the development of time-
dependent seismic fragility curves.” Journal of Structural Engineering, 136(12), 1497-
1511.

Ghosh, J., and Padgett, J. E. (2010). “Aging considerations in the development of time-
dependent seismic fragility curves.” Journal of Structural Engineering, 136(12), 1497-
1511.

Ghosh, J., and Padgett, J. E. (2011). “Probabilistic seismic loss assessment of aging bridges
using a component‐level cost estimation approach.” Earthquake Engineering &
Structural Dynamics, 40(15), 1743-1761.
Ghosh, J., and Sood, P. (2016). “Consideration of time-evolving capacity distributions and
improved degradation models for seismic fragility assessment of aging highway
bridges.” Reliability Engineering & System Safety, 154, 197-218.

224
Ghosh, J., Caprani, C. C., and Padgett, J. E. (2014). “Influence of traffic loading on the seismic
reliability assessment of highway bridge structures.” Journal of Bridge
Engineering, 19(3), 04013009.

Ghosh, J., Padgett, J. E., and Sánchez-Silva, M. (2015). “Seismic damage accumulation in
highway bridges in earthquake-prone regions.” Earthquake Spectra, 31(1), 115-135.
Ghosh, J., Rokneddin, K., Padgett, J. E., and Dueñas-Osorio, L. (2014). “Seismic reliability
assessment of aging highway bridge networks with field instrumentation data and
correlated failures, I: Methodology.” Earthquake Spectra, 30(2), 795-817.

Gidaris, I., Padgett, J. E., Barbosa, A. R., Chen, S., Cox, D., Webb, B., and Cerato, A. (2017).
“Multiple-hazard fragility and restoration models of highway bridges for regional risk
and resilience assessment in the United States: State-of-the-art review.” Journal of
structural engineering, 143(3), 04016188.

Goda, K., Kiyota, T., Pokhrel, R. M., Chiaro, G., Katagiri, T., Sharma, K., & Wilkinson, S.
(2015). “The 2015 Gorkha Nepal earthquake: insights from earthquake damage
survey.” Frontiers in Built Environment, 1, 8.

Griggs, G. B., Patsch, K., and Savoy, L. E. (2005). Living with the changing California coast.
Univ of California Press.

Gulerce, Z., Erduran, E., Kunnath, S. K., and Abrahamson, N. A. (2012). “Seismic demand
models for probabilistic risk analysis of near fault vertical ground motion effects on
ordinary highway bridges.” Earthquake engineering & structural dynamics, 41(2),
159-175.

Guo, X., and Chen, Z. (2016). “Lifecycle multihazard framework for assessing flood scour and
earthquake effects on bridge failure.” ASCE-ASME Journal of Risk and Uncertainty in
Engineering Systems, Part A: Civil Engineering, 2(2), C4015004.
Haghani, R., Al-Emrani, M., and Heshmati, M. (2012). “Fatigue-prone details in steel
bridges.” Buildings, 2(4), 456-476.

Hajializadeh, D., Stewart, M. G., Enright, B., and OBrien, E. (2016). “Spatial time-dependent
reliability analysis of reinforced concrete slab bridges subject to realistic traffic
loading.” Structure and Infrastructure Engineering, 12(9), 1137-1152.

225
Hamlet, A. F., and Lettenmaier, D. P. (1999). “Effects of climate change on hydrology and
water resources in the Columbia River Basin”. JAWRA Journal of the American Water
Resources Association, 35(6), 1597-1623.
Han, J., Song, Y., Wang, L., and Song, S. (2015). “Residual strain analysis of non-prestressed
reinforcement in PPC beams under fatigue loading.” Materials and Structures, 48,
1785-1802.

Han, X., and Frangopol, D. M. (2023). “Life-Cycle Risk-Based Optimal Maintenance Strategy
for Bridge Networks Subjected to Corrosion and Seismic Hazards.” Journal of Bridge
Engineering, 28(1), 04022128.

Han, Z., Ye, A., and Fan, L. (2010). “Effects of riverbed scour on seismic performance of high-
rise pile cap foundation.” Earthquake Engineering and Engineering Vibration, 9(4),
533.
Henry, D., and Ramirez-Marquez, J. E. (2012). “Generic metrics and quantitative approaches
for system resilience as a function of time.” Reliability Engineering & System
Safety, 99, 114-122.

Higgins, C., Lee, A. Y., Potisuk, T., and Forrest, R. W. (2007). “High-cycle fatigue of
diagonally cracked RC bridge girders: laboratory tests.” Journal of Bridge
Engineering, 12(2), 226-236.

Ikpong, A., and Bagchi, A. (2015). “New method for climate change resilience rating of
highway bridges.” Journal of Cold Regions Engineering, 29(3), 04014013.

IPCC (Intergovernmental Panel on Climate Change) (2007). The physical science basis.” In
Proc., Contribution of Working Group I to the 4th Assessment Report of the
Intergovernmental Panel on Climate Change, 996. Geneva: IPCC.
IRC-112 (2011). Code of practice for concrete road bridges. Indian Roads Congress;
New Delhi.
ISO (International Organization for Standardization) (2016). Mechanical vibration—Road surface
profiles—Reporting of measured data. ISO 8608. Geneva: ISO.

Kallias, A. N., and Imam, B. (2016). “Probabilistic assessment of local scour in bridge piers
under changing environmental conditions.” Structure and Infrastructure
Engineering, 12(9), 1228-1241.

226
Kameshwar, S., and Padgett, J. E. (2014). “Multi-hazard risk assessment of highway bridges
subjected to earthquake and hurricane hazards.” Engineering Structures, 78, 154-166.

Kameshwar, S., and Padgett, J. E. (2018). “Parameterized fragility assessment of bridges


subjected to pier scour and vehicular loads.” Journal of Bridge Engineering, 23(7),
04018044.

Kameshwar, S., and Padgett, J. E. (2018a). “Effect of vehicle bridge interaction on seismic
response and fragility of bridges.” Earthquake Engineering & Structural
Dynamics, 47(3), 697-713.

Kameshwar, S., and Padgett, J. E. (2018b). “Parameterized fragility assessment of bridges


subjected to pier scour and vehicular loads.” Journal of Bridge Engineering, 23(7),
04018044.

Kameshwar, S., and Padgett, J. E. (2018c). “Response and fragility assessment of bridge
columns subjected to barge-bridge collision and scour.” Engineering Structures, 168,
308-319.

Kaplan, H., and Seireg, A. (2002). “A base isolation system for bridges subjected to seismic
disturbances.” Earthquake engineering & structural dynamics, 31(5), 1093-1112.

Karamlou, A., and Bocchini, P. (2015). “Computation of bridge seismic fragility by large‐scale
simulation for probabilistic resilience analysis.” Earthquake Engineering & Structural
Dynamics, 44(12), 1959-1978.

Karuppusamy, B., Leo George, S., Anusuya, K., Venkatesh, R., Thilagaraj, P., Gnanappazham,
L., ... and Balabaskaran Nina, P. (2021). “Revealing the socio-economic vulnerability
and multi-hazard risks at micro-administrative units in the coastal plains of Tamil Nadu,
India.” Geomatics, Natural Hazards and Risk, 12(1), 605-630.

Kashani, M. M., Crewe, A. J., and Alexander, N. A. (2013). “Nonlinear stress–strain behaviour
of corrosion-damaged reinforcing bars including inelastic buckling.” Engineering
Structures, 48, 417-429.

Kassir, M. K., and Ghosn, M. (2002). “Chloride-induced corrosion of reinforced concrete


bridge decks.” Cement and Concrete Research, 32(1), 139-143.

Khakzand, L., Jalali, R. S., and Veisi, M. (2022). “Near‐fault ground motion effects on
pounding and unseating using an example of a three‐span, simply supported
bridge.” Earthquake Engineering and Resilience, 1(2), 164-195.

227
Khandel, O., and Soliman, M. (2019). “Integrated framework for quantifying the effect of
climate change on the risk of bridge failure due to floods and flood-induced
scour.” Journal of bridge engineering, 24(9), 04019090.

Khelifa, A., Garrow, L. A., Higgins, M. J., and Meyer, M. D. (2013). “Impacts of climate
change on scour-vulnerable bridges: Assessment based on HYRISK.” Journal of
infrastructure systems, 19(2), 138-146.

Kim, S. H., and Feng, M. Q. (2003). “Fragility analysis of bridges under ground motion with
spatial variation.” International Journal of Non-Linear Mechanics, 38(5), 705-721.

Kim, S. J. (2008). Seismic assessment of RC structures considering vertical ground motion.


PhD Dissertation, University of Illinois at Urbana-Champaign.
Kim, S. J., Holub, C. J., and Elnashai, A. S. (2011). “Analytical assessment of the effect of
vertical earthquake motion on RC bridge piers.” Journal of Structural
Engineering, 137(2), 252-260.

Kim, Y. J., and Heffernan, P. J. (2008). “Fatigue behavior of externally strengthened concrete
beams with fiber-reinforced polymers: State of the art.” Journal of Composites for
Construction, 12(3), 246-256.

Klemeš, V. (1986). “Operational testing of hydrological simulation models.” Hydrological


sciences journal, 31(1), 13-24.

Konstantinidis, D., Kelly, J. M., and Makris, N. (2008). Experimental investigation on the
seismic response of bridge bearings (pp. 1-339). Earthquake Engineering Research
Center, University of California, Berkeley.

Kumar, R., Gardoni, P., and Sanchez‐Silva, M. (2009). “Effect of cumulative seismic damage
and corrosion on the life‐cycle cost of reinforced concrete bridges.” Earthquake
Engineering & Structural Dynamics, 38(7), 887-905.
Kunnath, S. K., Erduran, E., Chai, Y. H., and Yashinsky, M. (2008). “Effect of near-fault
vertical ground motions on seismic response of highway overcrossings.” Journal of
Bridge Engineering, 13(3), 282-290.

Kwon, K., Frangopol, D. M., and Soliman, M. (2012). “Probabilistic fatigue life estimation of
steel bridges by using a bilinear S-N approach.” Journal of Bridge Engineering, 17(1),
58-70.

228
Kwon, O. S., and Elnashai, A. S. (2010). “Fragility analysis of a highway over-crossing bridge
with consideration of soil–structure interactions.” Structure and Infrastructure
Engineering, 6(1-2), 159-178.

LaFave, J., Fahnestock, L., Foutch, D. A., Steelman, J., Revell, J., Filipov, E., and Hajjar, J. F.
(2013). Experimental investigation of the seismic response of bridge bearings. FHWA-
ICT-13-002.

Lagasse, P. F., L. W. Zevenbergen, J. D. Schall, and P. E. Clopper (2001). Bridge scour and stream instability
countermeasures. Hydraulic Engineering Circular No. 23, FHWA NHI 01-003. Washington, DC:
Federal Highway Administration.

Laman, J. A., and Nowak, A. S. (1996). “Fatigue-load models for girder bridges.” Journal of
Structural Engineering, 122(7), 726-733.

Lehman, D. E., and Moehle, J. P. (2000). Performance-based seismic design of reinforced


concrete bridge columns. In Twelfth World Earthquake Engineering Conference (pp.
215-223).

Li, L., Li, C. Q., and Mahmoodian, M. (2019). “Prediction of fatigue failure of steel beams
subjected to simultaneous corrosion and cyclic loading.” Structures (Vol. 19, pp. 386-
393). Elsevier.

Li, T., Lin, J., and Liu, J. (2020). “Analysis of time-dependent seismic fragility of the offshore
bridge under the action of scour and chloride ion corrosion.” Structures, Vol. 28, pp.
1785-1801).

Liang, X., Lettenmaier, D. P., Wood, E. F., and Burges, S. J. (1994). “A simple hydrologically
based model of land surface water and energy fluxes for general circulation
models.” Journal of Geophysical Research: Atmospheres, 99(D7), 14415-14428.
Lu, Q., Harvey, J., Le, T., Lea, J., Quinley, R., Redo, D., and Avis, J. (2002). Truck traffic
analysis using weigh-in-motion (WIM) data in California. Institute of Transportation
Studies, University of California, Berkeley.

Lu, Q., Zhang, Y., and Harvey, J. T. (2009). “Estimation of truck traffic inputs for mechanistic–
empirical pavement design in California.” Transportation research record, 2095(1),
62-72.

229
Lu, Y., Hajirasouliha, I., & Marshall, A. M. (2018). “Direct displacement-based seismic design
of flexible-base structures subjected to pulse-like ground motions.” Engineering
Structures, 168, 276-289.

Lu, Y., Tang, W., Li, S., and Tang, M. (2018). “Effects of simultaneous fatigue loading and
corrosion on the behavior of reinforced beams.” Construction and Building
Materials, 181, 85-93.

Lu, Y., Wang, R., Han, Q., Yu, X., and Yu, Z. (2022). “Experimental investigation on the
corrosion and corrosion fatigue behavior of butt weld with G20Mn5QT cast steel and
Q355D steel under dry–wet cycle.” Engineering Failure Analysis, 134, 105977.

Lund University (2018) Climate change impact on existing bridges: A s r c r engineer’s


perspective. Lund University.

Luo, Y., Zheng, H., Zhang, H., and Liu, Y. (2021). “Fatigue reliability evaluation of aging
prestressed concrete bridge accounting for stochastic traffic loading and resistance
degradation.” Advances in Structural Engineering, 24(13), 3021-3029.

Ma, Y., Wang, G., Su, X., Wang, L., and Zhang, J. (2018). “Experimental and modelling of
the flexural performance degradation of corroded RC beams under fatigue
load.” Construction and Building Materials, 191, 994-1003.

Ma, Y., Xiang, Y., Wang, L., Zhang, J., and Liu, Y. (2014). “Fatigue life prediction for aging
RC beams considering corrosive environments.” Engineering Structures, 79, 211-221.

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., and Stojadinović, B. (2007). “Performance‐based seismic bridge design for
damage and loss limit states.” Earthquake engineering & structural dynamics, 36(13),
1953-1971.

Mackie, K. R., Kucukvar, M., Tatari, O., and Elgamal, A. (2016). “Sustainability metrics for
performance-based seismic bridge response.” Journal of Structural
Engineering, 142(8), C4015001.

Mackie, K., and Stojadinović, B. (2001). “Probabilistic seismic demand model for California
highway bridges.” Journal of Bridge Engineering, 6(6), 468-481.

230
Madabhushi, S. P. G., and Haigh, S. K. (2005). “The Bhuj, India earthquake of 26th January
2001: a field report by EEFIT.” Earthquake Engineering Field Investigation Team,
Institution of Structural Engineers, London, UK.

Maekawa, K., Gebreyouhannes, E., Mishima, T., and An, X. (2006). “Three-dimensional
fatigue simulation of RC slabs under traveling wheel-type loads.” Journal of Advanced
Concrete Technology, 4(3), 445-457.

Mahmoud, H., and Cheng, G. (2017). “Framework for lifecycle cost assessment of steel
buildings under seismic and wind hazards.” Journal of Structural Engineering, 143(3),
04016186.

Mandal, P., Rastogi, B. K., Satyanarayana, H. V. S., and Kousalya, M. (2004). “Results from
local earthquake velocity tomography: implications toward the source process involved
in generating the 2001 Bhuj earthquake in the lower crust beneath Kachchh
(India).” Bulletin of the Seismological Society of America, 94(2), 633-649.
Mander, J. B., Priestley, M. J., and Park, R. (1988). “Theoretical stress-strain model for
confined concrete.” Journal of structural engineering, 114(8), 1804-1826.

Manson, S. S. (1953). Behavior of materials under conditions of thermal stress (Vol. 2933).
National Advisory Committee for Aeronautics.

Marin-Artieda, C. C., and Whittaker, A. S. (2010). “Theoretical studies of the XY-FP seismic
isolation bearing for bridges.” Journal of Bridge Engineering, 15(6), 631-638.

Markogiannaki, O. (2019). “Climate change and natural hazard risk assessment framework for
coastal cable-stayed bridges.” Frontiers in built environment, 5, 116.

Maurer, E. P., Wood, A. W., Adam, J. C., Lettenmaier, D. P., and Nijssen, B. (2002). “A long-
term hydrologically based dataset of land surface fluxes and states for the conterminous
United States.” Journal of climate, 15(22), 3237-3251.
MES (2022). A design of axel load survey report for Bagodra-Dhandhuka-Ranpur-Umrala
road on both LHS and RHS side of the section in Dhandhuka city in state of Gujarat.
Doc./Issue No. 1. Surat, India: Mattest Engineering Service; 2022. p. 1-39.
Minaie, E., and Moon, F. (2017). “Practical and simplified approach for quantifying bridge
resilience.” Journal of Infrastructure Systems, 23(4), 04017016.

Miner, M. A. (1945). “Cumulative damage in fatigue.” J. Appl. Mech. Sep 1945, 12(3): A159-
A164

231
Mondal, A., and Mujumdar, P. P. (2016). “Detection of change in flood return levels under
global warming.” Journal of Hydrologic Engineering, 21(8), 04016021.
Mondoro, A., Frangopol, D. M., and Liu, L. (2018). “Bridge adaptation and management under
climate change uncertainties: A review.” Natural Hazards Review, 19(1), 04017023.

Mori, A., Moss, P. J., Carr, A. J., and Cooke, N. (1997). “Behavior of laminated elastomeric
bearings.” Structural engineering and mechanics: An international journal, 5(4), 451-
469.

MoRTH (2022). Annual Report 2022-2023. Government of India, Ministry of Road Transport
and Highway https://morth.nic.in/annual-report

Mosher (1984). Load-transfer criteria for numerical analysis of axially loaded piles in sand.
Part II: Load pile capacity curves for steel and concrete piles. Technical Rep. K-84-1.
Vicksburg, MS: US Army Engineer Waterways Experiment Station, US Army Corps
of Engineers.

Moshirabadi, S., Soltani, M., and Maekawa, K. (2015). “Seismic interaction of underground
RC ducts and neighboring bridge piers in liquefiable soil foundation.” Acta
Geotechnica, 10, 761-780.
Murty, A. S. R., Gupta, U. C., and Radha, A. (1995). “A new approach to fatigue strength
distribution for fatigue reliability evaluation.” International journal of fatigue, 17(2),
85-89.

Muthukumar, S (2003) A contact element approach with hysteresis damping for the analysis and design of
pounding in bridges. Ph.D. thesis, Dept. of Civil and Environmental Engineering, Georgia Institute
of Technology.

Narayan, J. P., Sharma, M. L., and Kumar, A. (2002). “A seismological report on the 26
January 2001 Bhuj, India earthquake.” Seismological Research Letters, 73(3), 343-
355.
NBI (2015). National Bridge Inventory. Retrieved from
https://www.fhwa.dot.gov/bridge/nbi.cfm

NCHRP (2014). Strategic issues facing transportation, volume 2: Climate change, extreme weather events,
and the highway system. Rep. No. Project 20-83 (5). Washington, DC: NCHRP.

NDMP (2019). National Disaster Management Plan – 2019. Accessed: 18th January, 2022.
Retrieved from https://ndma.gov.in/sites/default/files/PDF/ndmp-2019.pdf

232
Newmark, N. M., Blume, J. A., and Kapur, K. K. (1973). “Seismic design spectra for nuclear
power plants.” Journal of the Power Division, 99(2), 287-303.

Ni Choine, M., Kashani, M., Lowes, L. N., O'Connor, A., Crewe, A. J., Alexander, N. A., and
Padgett, J. E. (2016). “Nonlinear dynamic analysis and seismic fragility assessment of
a corrosion damaged integral bridge.” International Journal of Structural
Integrity, 7(2).
Nielson, B. G. (2005). Analytical fragility curves for highway bridges in moderate seismic
zones. Georgia Institute of Technology.

NZ. (2016). New Zealand Bridge Design Manual (2nd ed.). Wellington, New Zealand: New
Zealand Transport Agency.

Okada, K., Okamura, H., and Sonoda, K. (1978). “Fatigue failure mechanism of reinforced
concrete bridge deck slabs.” Transportation research record, 664, 136-144.

Ou, Y. C., Fan, H. D., and Nguyen, N. D. (2013). “Long‐term seismic performance of
reinforced concrete bridges under steel reinforcement corrosion due to chloride
attack.” Earthquake Engineering & Structural Dynamics, 42(14), 2113-2127.
Padgett, J. E., and DesRoches, R. (2008). “Methodology for the development of analytical
fragility curves for retrofitted bridges.” Earthquake Engineering & Structural
Dynamics, 37(8), 1157-1174.
Padgett, J. E., Ghosh, J., and Dueñas-Osorio, L. (2013). “Effects of liquefiable soil and bridge
modelling parameters on the seismic reliability of critical structural
components.” Structure and Infrastructure Engineering, 9(1), 59-77.
Pandit, A. S., Joshi, M. N., Bhargava, P., Shaikh, I., Ayachit, G. N., Raj, S. R., ... and
Bagatharia, S. B. (2015). “A snapshot of microbial communities from the Kutch: one
of the largest salt deserts in the World.” Extremophiles, 19, 973-987.
Pang, Y., Wei, K., He, H., and Wang, W. (2022). “Assessment of lifetime seismic resilience of
a long-span cable-stayed bridge exposed to structural corrosion.” Soil Dynamics and
Earthquake Engineering, 157, 107275.

Papazoglou, A. J., and Elnashai, A. S. (1996). “Analytical and field evidence of the damaging
effect of vertical earthquake ground motion.” Earthquake Engineering & Structural
Dynamics, 25(10), 1109-1137.

233
Paraskeva, T. S., Dimitrakopoulos, E. G., and Zeng, Q. (2017). “Dynamic vehicle–bridge
interaction under simultaneous vertical earthquake excitation.” Bulletin of Earthquake
Engineering, 15, 71-95.

PEER (2023). PEER Ground motion database. https://ngawest2.berkeley.edu/site. Accessed


March 11, 2023.
Perdikaris, P. C., and Beim, S. (1988). “RC bridge decks under pulsating and moving
load.” Journal of Structural Engineering, 114(3), 591-607.

Pierce, D. W., Cayan, D. R., Das, T., Maurer, E. P., Miller, N. L., Bao, Y., ... and Tyree, M.
(2013). “The key role of heavy precipitation events in climate model disagreements of
future annual precipitation changes in California.” Journal of Climate, 26(16), 5879-
5896.

Pipinato, A., Pellegrino, C., and Modena, C. (2011). “Fatigue assessment of highway steel
bridges in presence of seismic loading.” Engineering Structures, 33(1), 202-209.

Priestley, M. N., Seible, F., and Calvi, G. M. (1996). Seismic design and retrofit of bridges.
John Wiley & Sons.

Priestley, M. N., Verma, R., and Xiao, Y. (1994). “Seismic shear strength of reinforced
concrete columns.” Journal of structural engineering, 120(8), 2310-2329.
Ramadan, F., Smerzini, C., Lanzano, G., and Pacor, F. (2021). “An empirical model for the
vertical‐to‐horizontal spectral ratios for Italy.” Earthquake Engineering & Structural
Dynamics, 50(15), 4121-4141.
Ranjkesh, S. H., Asadi, P., and Hamadani, A. Z. (2019). “Seismic collapse assessment of
deteriorating RC bridges under multiple hazards during their life-cycle.” Bulletin of
earthquake engineering, 17, 5045-5072.

Rao, A. S., Lepech, M. D., and Kiremidjian, A. (2017). “Development of time-dependent


fragility functions for deteriorating reinforced concrete bridge piers.” Structure and
Infrastructure Engineering, 13(1), 67-83.

Rastogi, B. K. (2001). “Ground deformation study of Mw 7.7 Bhuj earthquake of


2001.” Episodes Journal of International Geoscience, 24(3), 160-165.

Rathje, E. M., Abrahamson, N. A., and Bray, J. D. (1998). “Simplified frequency content
estimates of earthquake ground motions.” Journal of geotechnical and
geoenvironmental engineering, 124(2), 150-159.

234
Richardson, E.V. and Davis, S.R (1995). Evaluating scour at bridges, Rep. No. FHWA-IP-
90-017 (HEC-18), Federal Administration, U.S. Department of Transportation,
Washington D.C.

Rocha, M., and Brühwiler, E. (2012). “Prediction of fatigue life of reinforced concrete bridges
using fracture mechanics. “Proceedings bridge maintenance, safety, management,
resilience and sustainability, 1(CONF), 3755-3761.

Rokneddin, K., Ghosh, J., Dueñas-Osorio, L., and Padgett, J. E. (2013). “Bridge retrofit
prioritisation for ageing transportation networks subject to seismic hazards.” Structure
and Infrastructure Engineering, 9(10), 1050-1066.
Rokneddin, K., Ghosh, J., Dueñas-Osorio, L., and Padgett, J. E. (2014). “Seismic reliability
assessment of aging highway bridge networks with field instrumentation data and
correlated failures, II: Application.” Earthquake Spectra, 30(2), 819-843.
Romstad, K., Kutter, B., Maroney, B., Vanderbilt, E., Griggs, M., and Chai, Y. H. (1995).
Experimental measurements of bridge abutment behavior. Rep. No. UCD-STR-95, 1.

Roy, T., and Matsagar, V. (2021). “Multi-hazard analysis and design guidelines:
recommendations for structure and infrastructure systems in the Indian
context.” Current Science, 121(1), 44-55.

Saadeghvariri, M. A., and Foutch, D. A. (1991). “Dynamic behaviour of R/C highway bridges
under the combined effect of vertical and horizontal earthquake motions.” Earthquake
engineering & structural dynamics, 20(6), 535-549.

Sahana, M., Rehman, S., Ahmed, R., and Sajjad, H. (2021). “Assessing losses from multi-
hazard coastal events using Poisson regression: empirical evidence from Sundarban
Biosphere Reserve (SBR), India.” Journal of Coastal Conservation, 25, 1-10.

Saidi, A., Hamouine, A., Bousserhane, I. K., and Fali, L. (2017). “Study of vehicle-bridge
coupled vibration using matlab/simulink.” Int. J. Civ. Eng. Technol, 8(12), 502-511.

Sakano, M., and Wahab, M. A. (2001). “Extremely low cycle (ELC) fatigue cracking behaviour
in steel bridge rigid frame piers.” Journal of Materials Processing Technology, 118(1-
3), 36-39.

SDC (2019). Caltrans Seismic Design Criteria. Version 2.0, State of California Department of
Transportation, California.

235
Shamsabadi, A., and Yan, L. (2008). “Closed-form force-displacement backbone curves for
bridge abutment-backfill systems.” In Geotechnical earthquake engineering and soil
dynamics IV (pp. 1-10).

Sharma, N., Tabandeh, A., and Gardoni, P. (2018). “Resilience analysis: A mathematical
formulation to model resilience of engineering systems.” Sustainable and Resilient
Infrastructure, 3(2), 49-67.

Sheikh, S. A., and Yau, G. (2002). “Seismic behavior of concrete columns confined with steel
and fiber-reinforced polymers.” Structural Journal, 99(1), 72-80.

Shekhar, S., Ghosh, J., and Padgett, J. E. (2018). “Seismic life-cycle cost analysis of ageing
highway bridges under chloride exposure conditions: Modelling and
recommendations.” Structure and Infrastructure Engineering, 14(7), 941-966.
Shigefuji, M., Takai, N., Bijukchhen, S., Ichiyanagi, M., Rajaure, S., Dhital, M. R., ... and
Sasatani, T. (2022). “Strong ground motion data of the 2015 Gorkha Nepal earthquake
sequence in the Kathmandu Valley.” Scientific Data, 9(1), 513.

Shinozuka, M., Dong, X., Jin, X., and Cheng, T. C. (2005). “Seismic performance analysis for
the ladwp power system.” In 2005 IEEE/PES Transmission & Distribution Conference
& Exposition: Asia and Pacific (pp. 1-6). IEEE.

Shinozuka, M., Feng, M. Q., Lee, J., and Naganuma, T. (2000). “Statistical analysis of fragility
curves.” Journal of engineering mechanics, 126(12), 1224-1231.

Shrestha, B. (2015). “Seismic response of long span cable-stayed bridge to near-fault vertical
ground motions.” KSCE Journal of Civil Engineering, 19, 180-187.

Shrestha, S. (2014). “Probabilistic seismic hazard analysis of Kathmandu city,


Nepal.” International Journal of Engineering Research and General Science, 2(1), 24-
33.

Simon, J., Bracci, J. M., and Gardoni, P. (2010). “Seismic response and fragility of deteriorated
reinforced concrete bridges.” Journal of Structural Engineering, 136(10), 1273-1281.

Somerville, P., N. Smith, S. Punyamurthula, and J. Sun. (1997). Development of ground motion time
histories for Phase 2 of the FEMA/SAC Steel Project. SAC Joint Venture Project Rep. No.
SAC/BD-97/04. Richmond, CA: SAC Steel Project

236
Song, H. W., You, D. W., Byun, K. J., and Maekawa, K. (2002). “Finite element failure
analysis of reinforced concrete T-girder bridges.” Engineering Structures, 24(2), 151-
162.

Song, L., Fan, Z., and Hou, J. (2019). “Experimental and analytical investigation of the fatigue
flexural behavior of corroded reinforced concrete beams.” International Journal of
Concrete Structures and Materials, 13, 1-14.

Spellman, D. L., and Stratfull, R. F. (1969). Chlorides and bridge deck deterioration (No. D-
3-11). Division of Highways, Materials and Research Department.

Statisca (2023). Heavy truck production volume in India from 2011 to 2021. Statista. Published
April 07, 2022. Accessed February 22, 2023.
https://www.statista.com/statistics/261212/heavy-truck-production-in-india/
Steelman, J. S., Fahnestock, L. A., Filipov, E. T., LaFave, J. M., Hajjar, J. F., and Foutch, D.
A. (2013). “Shear and friction response of non seismic laminated elastomeric bridge
bearings subject to seismic demands.” Journal of Bridge Engineering, 18(7), 612-623.
Stefanidou, S. and Kappos A. (2017). “Methodology for the development of bridge-specific
fragility curves.” Earthquake Engineering and Structural Dynamics, vol. 46, pp 73-93.

Stein, S. M., Young, G. K., Trent, R. E., and Pearson, D. R. (1999). “Prioritizing scour
vulnerable bridges using risk.” Journal of infrastructure systems, 5(3), 95-101.

Stevens, V. L., Shrestha, S. N., and Maharajan, D. K. (2018). “Probabilistic seismic hazard
assessment of Nepal.” Bulletin of the Seismological Society of America, 108(6), 3488-
3510.

Stewart, M. G. (2004). “Spatial variability of pitting corrosion and its influence on structural
fragility and reliability of RC beams in flexure.” Structural safety, 26(4), 453-470.
Stewart, M. G., and Al-Harthy, A. (2008). “Pitting corrosion and structural reliability of
corroding RC structures: Experimental data and probabilistic analysis.” Reliability
engineering & system safety, 93(3), 373-382.
Stewart, M. G., and Rosowsky, D. V. (1998). “Time-dependent reliability of deteriorating
reinforced concrete bridge decks.” Structural safety, 20(1), 91-109.

Suaznabar, O., Huang, C., Xie, Z., Shen, J., Kerenyi, K., Bergendahl, B., and Kilgore, R.
(2017). Hydraulic performance of shallow foundations for the support of vertical-wall

237
bridge abutments (No. FHWA-HRT-17-013). United States. Federal Highway
Administration.

Sun, J., Huang, Q., and Ren, Y. (2015). “Performance deterioration of corroded RC beams and
reinforcing bars under repeated loading.” Construction and Building Materials, 96,
404-415.

Sung, Y. C., and Su, C. K. (2011). “Time-dependent seismic fragility curves on optimal
retrofitting of neutralised reinforced concrete bridges.” Structure and Infrastructure
Engineering, 7(10), 797-805.
Tapia-Hernández, E., Perea, T., and Islas-Mendoza, M. A. (2017). “Design assessment of
short-span steel bridges.” International Journal of Civil Engineering, 15, 319-332.

Thanapol, Y., Akiyama, M., and Frangopol, D. M. (2016). “Updating the seismic reliability of
existing RC structures in a marine environment by incorporating the spatial steel
corrosion distribution: Application to bridge piers.” Journal of Bridge
Engineering, 21(7), 04016031.
Thapa, S., Shrestha, Y., and Gautam, D. (2022, March). “Seismic fragility analysis of RC
bridges in high seismic regions under horizontal and simultaneous horizontal and
vertical excitations.” In Structures (Vol. 37, pp. 284-294). Elsevier.

US Dept. of Labor (2018). U.S. Department of Labor. Accessed October 26, 2018.
https://www.dol.gov

USBR (2016). Climate change adaptation strategy: 2016 progress report. Washington, DC:
USBR.

USGS (2013). National seismic hazard mapping project. http://geohazards.usgs.gov/hazardtool/i (Dec. 18,
2017).

USGS (2019). Guidelines for determining flood flow frequency—Bulletin17C. Rep. No. 4-B5.
Washington, DC: USGS

Val, D. V., and Melchers, R. E. (1997). “Reliability of deteriorating RC slab bridges.” Journal
of structural engineering, 123(12), 1638-1644.
Val, D. V., Stewart, M. G., and Melchers, R. E. (2000). “Life‐cycle performance of RC bridges:
probabilistic approach.” Comp er‐Aided Civi nd Infr s r c re Engineering, 15(1),
14-25.

238
Veletzos, M. J., Restrepo, J. I., and Seible, F. (2006). Seismic response of precast segmental
bridge superstructures (No. UCSD/SSRP-06/18). California. Dept. of Transportation.

Veletzos, M. J., Restrepo, J. I., and Seible, F. (2006). Seismic response of precast segmental
bridge superstructures (No. UCSD/SSRP-06/18). California. Dept. of Transportation.

Venkittaraman, A., and Banerjee, S. (2014). “Enhancing resilience of highway bridges through
seismic retrofit.” Earthquake Engineering & Structural Dynamics, 43(8), 1173-1191.

Vishwanath, B. S., and Banerjee, S. (2019). “Life-cycle resilience of aging bridges under
earthquakes.” Journal of Bridge Engineering, 24(11), 04019106.

Vu, K. A. T., and Stewart, M. G. (2000). “Structural reliability of concrete bridges including
improved chloride-induced corrosion models.” Structural safety, 22(4), 313-333.

Vu, K. A. T., and Stewart, M. G. (2000). “Structural reliability of concrete bridges including
improved chloride-induced corrosion models.” Structural safety, 22(4), 313-333.
Wan, H. P., Su, L., Frangopol, D. M., Chang, Z., Ren, W. X., and Ling, X. (2021). “Seismic
Response of a Bridge Crossing a Canyon to Near-Fault Acceleration-Pulse Ground
Motions.” Journal of Bridge Engineering, 26(6), 05021006.

Wang, S. C., Liu, K. Y., Chen, C. H., and Chang, K. C. (2015). “Experimental investigation
on seismic behavior of scoured bridge pier with pile foundation.” Earthquake
engineering & Structural dynamics, 44(6), 849-864.
Wang, W., Deng, L., and Shao, X. (2016). “Number of stress cycles for fatigue design of
simply-supported steel I-girder bridges considering the dynamic effect of vehicle
loading.” Engineering Structures, 110, 70-78.

Wang, Z., Dueñas‐Osorio, L., and Padgett, J. E. (2013). “Seismic response of a bridge–soil–
foundation system under the combined effect of vertical and horizontal ground
motions.” Earthquake Engineering & Structural Dynamics, 42(4), 545-564.

Wang, Z., Dueñas-Osorio, L., and Padgett, J. E. (2014). “Influence of scour effects on the
seismic response of reinforced concrete bridges.” Engineering structures, 76, 202-214.

Wang, Z., Dueñas-Osorio, L., and Padgett, J. E. (2014b). “Influence of soil-structure


interaction and liquefaction on the isolation efficiency of a typical multispan continuous
steel girder bridge.” Journal of Bridge Engineering, 19(8), A4014001.

239
Wang, Z., Padgett, J. E., and Dueñas-Osorio, L. (2013). “Influence of vertical ground motions
on the seismic fragility modeling of a bridge-soil-foundation system.” Earthquake
Spectra, 29(3), 937-962.

Wang, Z., Padgett, J. E., and Dueñas-Osorio, L. (2014). “Risk-consistent calibration of load
factors for the design of reinforced concrete bridges under the combined effects of
earthquake and scour hazards.” Engineering Structures, 79, 86-95.

Warn, G. P., and Whittaker, A. S. (2008). “Vertical earthquake loads on seismic isolation
systems in bridges.” Journal of structural engineering, 134(11), 1696-1704.

Warn, G. P., Whittaker, A. S., and Constantinou, M. C. (2007). “Vertical stiffness of


elastomeric and lead–rubber seismic isolation bearings.” Journal of structural
engineering, 133(9), 1227-1236.
Welty, J., and Zeng, X. (2021). “Characteristics and causes of extreme snowmelt over the
conterminous United States.” Bulletin of the American Meteorological Society, 102(8),
E1526-E1542.

Wen, Y. K. (2001). “Minimum lifecycle cost design under multiple hazards.” Reliability
Engineering & System Safety, 73(3), 223-231.

Weyers, R. E., Fitch, M. G., Larsen, E. P., Al-Qadi, I. L., Chamberlin, W. P., and Hoffman, P.
C. (1994). Concrete bridge protection and rehabilitation: Chemical and physical
techniques. Service life estimates (No. SHRP-S-668).

Wibowo, H., and Sritharan, S. (2022). “Effects of vertical ground acceleration on the seismic
moment demand of bridge superstructure connections.” Engineering Structures, 253,
113820.

Wilson, P., and Elgamal, A. (2008). “Full scale bridge abutment passive earth pressure tests
and calibrated models.” In Proceedings of the 14th world conference on earthquake
engineering, Beijing, China.

Wilson, T., Chen, S., and Mahmoud, H. (2015). “Analytical case study on the seismic
performance of a curved and skewed reinforced concrete bridge under vertical ground
motion.” Engineering structures, 100, 128-136.

Wood, A. W., Leung, L. R., Sridhar, V., and Lettenmaier, D. P. (2004). “Hydrologic
implications of dynamical and statistical approaches to downscaling climate model
outputs.” Climatic change, 62(1-3), 189-216.

240
Xiang, N., and Li, J. (2017). “Experimental and numerical study on seismic sliding mechanism
of laminated-rubber bearings.” Engineering Structures, 141, 159-174.

Xiang, N., and Li, J. (2018). “Effect of exterior concrete shear keys on the seismic performance
of laminated rubber bearing-supported highway bridges in China.” Soil Dynamics and
Earthquake Engineering, 112, 185-197.
Yan, D., Luo, Y., Lu, N., Yuan, M., and Beer, M. (2017). “Fatigue stress spectra and reliability
evaluation of short-to medium-span bridges under stochastic and dynamic traffic
loads.” Journal of Bridge Engineering, 22(12), 04017102.

Yan, D., Luo, Y., Lu, N., Yuan, M., and Beer, M. (2017). “Fatigue stress spectra and reliability
evaluation of short-to medium-span bridges under stochastic and dynamic traffic
loads.” Journal of Bridge Engineering, 22(12), 04017102.

Yang, D. Y., and Frangopol, D. M. (2019). “Physics-based assessment of climate change


impact on long-term regional bridge scour risk using hydrologic modeling: Application
to Lehigh river watershed.” Journal of Bridge Engineering, 24(11), 04019099.

Yaohan, D., Liu, H., Liu, Y., Song, J., Tan, K., Li, L., ... and Wang, Q. (2022). “Strength
degradation of a ferrite-bainite weathering steel subjected to corrosion fatigue
environment.” Corrosion Science, 208, 110690.

Yen, B. T., Hodgson, I. C., Zhou, Y. E., and Crudele, B. B. (2013). “Bilinear S-N Curves and
Equivalent Stress Ranges for Fatigue Life Estimation.” Journal of Bridge
Engineering, 18(1), 26-30.

Yilmaz, T. (2015). Risk Assessment of Highway Bridges Under Multi-Hazard Effect of Flood
Induced Scour and Earthquake. PhD Dissertation, The Pennsylvania State University.

Yilmaz, T., Banerjee, S., and Johnson, P. A. (2016). “Performance of two real-life California
bridges under regional natural hazards.” Journal of Bridge Engineering, 21(3),
04015063.

Yilmaz, T., Banerjee, S., and Johnson, P. A. (2018). “Uncertainty in risk of highway bridges
assessed for integrated seismic and flood hazards.” Structure and Infrastructure
Engineering, 14(9), 1182-1196.
Yohe, G., and Strzepek, K. (2007). “Adaptation and mitigation as complementary tools for
reducing the risk of climate impacts.” Mitigation and adaptation strategies for global
change, 12, 727-739.

241
Zhang, J., Huo, Y., Brandenberg, S. J., and Kashighandi, P. (2008). “Effects of structural
characterizations on fragility functions of bridges subject to seismic shaking and lateral
spreading.” Earthquake Engineering and Engineering Vibration, 7, 369-382.
Zhang, Q., and Alam, M. S. (2019). “Performance-based seismic design of bridges: a global
perspective and critical review of past, present and future directions.” Structure and
Infrastructure Engineering, 15(4), 539-554.

Zhang, T., Zhang, X., Li, P., Li, H., Li, X., and Zou, Y. (2023). “Experimental Research on
Fatigue Performance of Reinforced Concrete T-Shaped Beams under Corrosion–
Fatigue Coupling Action.” Materials, 16(3), 1257.

Zhang, W., Song, X., Gu, X., and Li, S. (2012). “Tensile and fatigue behavior of corroded
rebars.” Construction and Building Materials, 34, 409-417.

Zhang, W., Ye, Z., Gu, X., Liu, X., and Li, S. (2017). “Assessment of fatigue life for corroded
reinforced concrete beams under uniaxial bending.” Journal of Structural
Engineering, 143(7), 04017048.

Zhao, Y. X., Gao, Q., and Wang, J. N. (2000). “An approach for determining an appropriate
assumed distribution of fatigue life under limited data.” Reliability Engineering &
System Safety, 67(1), 1-7.

Zheng, Y., Dong, Y., and Li, Y. (2018). “Resilience and life-cycle performance of smart
bridges with shape memory alloy (SMA)-cable-based bearings.” Construction and
Building Materials, 158, 389-400.

Zhong, J., Gardoni, P., and Rosowsky, D. (2012). “Seismic fragility estimates for corroding
reinforced concrete bridges.” Structure and Infrastructure Engineering, 8(1), 55-69.
Zhong, J., Jeon, J. S., Shao, Y. H., and Chen, L. (2019). “Optimal intensity measures in
probabilistic seismic demand models of cable-stayed bridges subjected to pulse-like
ground motions.” Journal of Bridge Engineering, 24(2), 04018118.
Zhong, J., Mao, Y., and Yuan, X. (2023). “Lifetime seismic risk assessment of bridges with
construction and aging considerations.” Structures (Vol. 47, pp. 2259-2272). Elsevier.

Zhou, Y., Banerjee, S., and Shinozuka, M. (2010). “Socio-economic effect of seismic retrofit
of bridges for highway transportation networks: a pilot study.” Structure and
Infrastructure Engineering, 6(1-2), 145-157.

242
Zhu, B., and Frangopol, D. M. (2013). “Risk-based approach for optimum maintenance of
bridges under traffic and earthquake loads.” Journal of structural engineering, 139(3),
422-434.

Zhu, B., and Frangopol, D. M. (2016a). “Time-variant risk assessment of bridges with partially
and fully closed lanes due to traffic loading and scour.” Journal of Bridge
Engineering, 21(6), 04016021.

243

You might also like