You are on page 1of 21

Bulletin of Earthquake Engineering

https://doi.org/10.1007/s10518-019-00621-4

ORIGINAL RESEARCH

Seismic fragility curves for California concrete bridges


with flared two‑column bents

Jong‑Su Jeon1   · Sujith Mangalathu2   · Sang‑Youl Lee3 

Received: 10 August 2018 / Accepted: 13 April 2019


© Springer Nature B.V. 2019

Abstract
The increased aesthetic appeal of columns with flares resulted in bridge configurations
with flared columns during the 1970–1990 design periods in California. Earthquake recon-
naissance reports indicated that the flare is detrimental on the structural functionality of the
bridge with some geometric configurations. This research examines the relative vulnerabil-
ity of bridge configurations with flared and non-flared columns for three-span two-column
bent bridges with oblong cross-sections. The numerical model of the columns with and
without flare is calibrated with existing experimental results. The calibrated models are
used to develop the fragility curves for California bridges. Fragility curves for each bridge
class are generated by accounting for the material, geometric, and structural uncertainties,
and the effects of column flare on the seismic demand and fragility of bridges are investi-
gated in detail in this paper. Results reveal that the presence of column flare for the case
study bridges increases the stiffness and strength of columns because of a low potential of
column shear failure, resulting in the reduction of bridge vulnerability and have beneficial
effects on the structural functionality. Thus, the flare does not necessarily reduce the seis-
mic performance for all bridge cases and depends on the failure mode associated with the
bridge configuration. The fragility curves suggested in this research can be used by stake-
holders in deciding the retrofitting and maintenance strategies of bridges with architectural
flares.

Keywords  Flared columns · Fragility curves · Model validation · Oblong column cross-
section

* Sujith Mangalathu
sujithmangalath@ucla.edu
Jong‑Su Jeon
jongsujeon@hanyang.ac.kr
Sang‑Youl Lee
lsy@anu.ac.kr
1
Department of Civil and Environmental Engineering, Hanyang University, Seoul 04763,
Republic of Korea
2
Department of Civil and Environmental Engineering, University of California, Los Angeles,
CA 90095, USA
3
Department of Civil Engineering, Andong National University, Andong,
Gyeongsangbuk‑do 36729, Republic of Korea

13
Vol.:(0123456789)
Bulletin of Earthquake Engineering

1 Introduction

Past earthquakes have demonstrated that highway bridges, as key components of trans-
portation networks, are one of the most vulnerable components. The damage state of the
bridge after an earthquake is critical in deciding the emergency or ordinary traffic. The
likelihood of bridge damage during a seismic event can be obtained through fragility
curves. Fragility curves are conditional probability statements that give the likelihood that
a structure or component of that structure will meet or exceed a certain level of damage for
a given ground motion intensity measure (IM).
Fragility curves are deployed these days by emergency management agencies to assess
the likelihood of bridge damage and to decide the operational state of the bridge (emer-
gency or ordinary traffic) following an earthquake (Mangalathu 2017). The significance
of bridge fragility curves leads to a number of studies on bridge fragilities (Mackie and
Stojadinovic 2001; Nielson and DesRoches 2007; Pan et al. 2007; Banerjee and Shinozuka
2008; Seo and Linzell 2012; Huo and Zhang 2013; Ramanathan 2012; Dimitrakopoulos
and Paraskeva 2015; Monteiro et  al. 2016; Jeon et  al. 2017; Rogers and Seo 2017; Seo
and Rogers 2017; Mangalathu et al. 2016, 2017a, 2018a). Insights from these studies are
valuable in identifying the seismic vulnerability, retrofitting strategies, seismic resilience,
and post-earthquake management decisions. Although initial studies (Mackie and Stojadi-
novic 2001; Nielson and DesRoches 2007; Pan et al. 2007; Banerjee and Shinozuka 2008,
Huo and Zhang 2013) were focused on bridge-specific fragility curves, recent studies are
directed towards the generation of fragility curves for bridge classes (Mangalathu 2017;
Ramanathan 2012; Mangalathu et al. 2016). However, none of the aforementioned studies
accounted for the effect of column flare on the generation of seismic fragilities.
The increased aesthetic appeal of flared bridge columns inspired the construction of
flared columns during the years between the 1970 and 1990. Especially, one-way flared
columns are commonly used to provide more support to the cap beam under eccentric live
loads. Most of the flared segments are minimally reinforced and are assumed to have a
negligible contribution to the seismic design strength of the columns (Saiidi et al. 2001).
However, the 1994 Northridge earthquake revealed some seismic performance problems
inherent in flared columns such as column shear failure due to the reduction of shear span
resulting from the shift of plastic hinge, which motivated some experimental and analytical
studies on flared bridge columns (Sanchez et al. 1997; Nada et al. 2003). These experimen-
tal studies concluded that architectural flares with minimum longitudinal reinforcement
behave essentially the same as structural flares and increase the shear and flexural demand
of the columns. The recommendation from these studies prompted California Department
of Transportation (Caltrans) to have a retrofit and design strategy to separate the flare from
the bent cap using a gap between the top of the flare and the beam bottom surface (Caltrans
2006). Caltrans recommends that the minimum gap thickness shall be 102 mm to mitigate
undesirable failure from the existence of flares.
Although the experimental and analytical studies have identified the importance of
flare in dictating the seismic demand and failure mode, very few studies (Soleimani
et al. 2017) accounted for the effect of flares on the bridge fragilities. Soleimani et al.
(2017) noted that the flare on circular columns can increase the seismic vulnerability
of bridges, and the increase in the median value of fragility varies from 0.75 to 15.89%
depending on the limit state under consideration. Per Mangalathu (2017), more than
30% of California bridges constructed during 1970–1990 are with oblong column (dou-
ble interlocking-spiral) cross-sections, as shown in Fig. 1. Experimental (Correal et al.

13
Bulletin of Earthquake Engineering

Fig. 1  Configuration and component models of selected bridge classes

2007; Tanaka and Park 1993) and analytical (Mangalathu et al. 2017b) studies revealed
that the oblong column cross-sections have a distinct seismic behavior compared to cir-
cular columns and the oblong columns are less vulnerable than the circular columns.
This paper investigates the effect of column flares and the flare gap on columns with
oblong cross-sections for the first time. Three-span integral and seat abutment bridges
with two-column bents, which are common in California (Mangalathu 2017), are cho-
sen as a case study in this paper to demonstrate the flare effects on seismic fragilities.
The analytical model developed in OpenSees (McKenna 2011) is initially calibrated
with experimental results. The subsequent section investigates the effect of flare design
details and flare gap on the seismic vulnerability of the case study bridge class. The sali-
ent observations noted from this research are finally provided.

13
Bulletin of Earthquake Engineering

2 Fragility curve development for selected bridge classes

2.1 Fragility curve development

The multi-phase fragility framework used by previous studies on bridge fragilities is used
in this research to generate fragility curves of bridge classes (Mangalathu 2017; Nielson
and DesRoches 2007; Ramanathan 2012). In this framework, input variables are sampled
across the range of uncertain parameters to generate statistically significant yet nominally
identical three–dimensional bridge models. The bridge models are randomly paired with
the selected suite of ground motions. Nonlinear dynamic analyses are performed for all
bridge-ground motion pairs to obtain the maximum response of various bridge compo-
nents. The probabilistic seismic demand models (PSDMs) are then estimated through a
linear regression on the demand parameters and intensity measures (IM) in a logarithmic
space. The convolution of the PSDMs with the capacity models of each bridge component
yields the component fragility curves. Assuming lognormal distribution for D and C, the
component fragility can be obtained using Eq. (1) (Cornell et al. 2002)

⎡ � � ⎤
⎢ ln SD ∕SC ⎥
P[D > C�IM] = Φ⎢ � ⎥ (1)
⎢ 𝛽D�IM
2
+ 𝛽C2 ⎥
⎣ ⎦

where SD is the median estimate of the demand as a function of the IM, SC is the median
estimate of the capacity, βD|IM is the dispersion of the demand conditioned on the IM, βC
is the dispersion of the capacity, and Φ[·] is the standard normal cumulative distribution
function. SD and βD|IM can be estimated from the PSDMs. The components fragility curves
can be integrated into system fragility curves through the generation of joint probabilistic
seismic demand models and Monte Carlo simulation following the work of Nielson and
DesRoches (2007).

2.2 Bridge class characteristics for fragility analysis

To investigate the effect of column flare, three-span integral and seat abutment bridges
with two-column bents and oblong cross-sections are used in this research. The integral
abutments are cast monolithic with the superstructure. Because the integral abutments
engage the backfill soil during seismic action, they provide a good source of energy dis-
sipation and reduce the potential of span unseating. Seat abutments provide a bearing sup-
port to the superstructure, which is restrained longitudinally by the abutment backwall
and transversely by the piles and the shear key. The stiffness and resistance to the seismic
action increase when the deck is in contact with the abutment backwall in the longitudinal
direction. Three-span bridges prevail more than 25% of California bridge inventory (Ram-
anathan 2012) and around 25% of multi-column bent bridges in California constructed
after 1970 are with oblong cross-sections (Mangalathu 2017). In California, multi-column
bents have a pinned connection at the column base while single column bents are designed
to have a fixed connection at the base (Priestley et  al. 1996). A typical column-footing
connection for a pinned column is shown in Fig.  1. Under the assumption of the pinned
connection, there is no restraint against column rotation at the column base, and thus the

13
Bulletin of Earthquake Engineering

column essentially behaves as a cantilever in both directions. Also, because of the moment-
resisting connection between the superstructure and the column top, the entire energy dis-
sipation happens at the top of the column by forming a plastic hinge at the column top.
This research selects a constant column cross-section as a baseline (prismatic) column,
which is the most common for two-column bent bridges from the extensive plan reviews
(Mangalathu 2017); the prismatic column used here has a rectangular section with 914 mm
by 1524 mm, as shown in Fig. 1. For the flared columns, the parabolic shape of one-way
and two-way flares is determined per the recommendation of Caltrans (2006), as illustrated
in Fig. 2. For the one-way flared column, the height of the parabolic flare is 2743 mm and
the column width at the column top is 2438 mm. In the case of the two-way flared column,
the front view is the same as that for the one-way flared column while for its side view, the
depth of the column at the soffit is 1372 mm.
The modeling parameters used to characterize the structural, geometric, and material
uncertainties are indicated in Table 1. The table presents the mean value and standard devi-
ation and associated distribution for each parameter. The parameters are collected based on
modeling assumptions of bridge components. These parameters are obtained through the
manual plan reviews of more than 1000 bridges in California (Mangalathu 2017) and can
reflect the bridge inventory in California. To avoid the unlikely combinations in generat-
ing the bridge model for dynamic analyses, the parameters are truncated at two times the
standard deviation from the mean value.

2.3 Ground motions and capacity models

As part of the PEER transportation system research program, Baker et al. (2011) sug-
gested 160 ground motions for the risk assessment of infrastructural systems in Cali-
fornia. The ground motions are scaled by a factor of two to have the sufficient response
data for IMs higher than the Palmdale response spectrum which corresponds to a hazard

Fig. 2  Geometry of one-way and two-way flare columns

13
Bulletin of Earthquake Engineering

Table 1  Uncertainty parameters of bridges and their probability distribution (Mangalathu 2017)


Parameter Type Parameters Truncated limit
Mean SD Lower Upper

Superstructure (pre-stressed concrete)


 Main-span length, Lm (m) N 47.2 13.72 20.4 74.1
 Ratio of side-span to main-span length, (η = Ls/Lm) N 0.75 0.2 0.4 1.0
 Deck width, Dw (m) (five-cell deck) N 17.4 2.44 13.4 21.4
Interior bent
 Concrete compressive strength, fc (MPa) N 31.4 3.86 24.0 39.0
 Rebar yield strength, fy (Mpa) N 476.0 37.92 400.0 550.0
 Column clear height, Hc (m) LN 7.1 1.15 4.9 9.4
 Column longitudinal reinforcement ratio, ρl U 0.02 0.006 0.01 0.03
 Column transverse reinforcement ratio, ρt U 0.009 0.003 0.004 0.013
Deep foundation (pile group)
 Translational stiffness, Kft (kN/mm) LN 175.0 23.4 135.0 227.0
Exterior bent (seat abutment on piles)
 Abutment backwall height, Ha (m) LN 3.39 0.69 2.2 4.9
 Pile stiffness, Kp (kN/mm) LN 0.093 0.033 0.044 0.174
 Backfill type, BT (sand vs. clay) B – – – –
Bearing (elastomeric bearing)
 Stiffness per deck width, Kb (N/mm/mm) LN 0.63 0.30 0.24 1.40
 Coefficient of friction of pad, μb N 0.3 0.1 0.1 0.5
Gap
 Longitudinal (pounding), Δl (mm) LN 23.3 12.4 7.6 55.4
 Transverse (shear key), Δt (mm) U 19.1 11.0 0 38.1
Other parameters
 Mass factor, mf U 1.05 0.06 0.95 1.15
 Damping, ξ N 0.045 0.0125 0.02 0.07
 Acceleration for shear key capacity (g), ask LN 1 0.2 0.8 1.2
 Earthquake direction (fault normal FN vs. parallel FP), ED B – – – –

N normal, LN lognormal, U uniform, B Bernoulli distribution

level of 10% probability of exceedance in 100  years (Ramanathan 2012). Thus, the
expanded suite of 320 ground motions is used in this research for the fragility curve
development. For fragility analysis, six engineering demand parameters (EDPs) are
selected and recorded during dynamic analyses: maximum column curvature ductility
(COL), maximum passive, active, and transverse abutment deformation (ABP, ABA,
and ABT in mm), maximum bearing deformation (BRG in mm), and maximum unseat-
ing deformation (UST in mm). For the selected bridge classes (constructed between
1970 and 1990), the capacity-based limit state models associated with the above EDPs
are presented in Table 2. These models are based on the statistical analysis of extensive
experimental results (Mangalathu 2017). Following the work of Mangalathu (2017), the
dispersion is assigned as 0.35 in a subjective manner due to lack of sufficient informa-
tion and is adopted as the same across the components and the respective limit states.

13
Bulletin of Earthquake Engineering

Table 2  Limit state models of various bridge components (Mangalathu 2017)


Component Median value, Sc Dispersion, (βc)
Slight ­(LS1) Moderate ­(LS2) Extensive ­(LS3) Complete ­(LS4)

Column curvature ductility (COL) 1 5 8 11 0.35


Passive abutment response (ABP, mm) 76 254 – – 0.35
Active abutment response (ABA, mm) 38 102 – – 0.35
Transverse abutment response (ABT, mm) 25 102 – – 0.35
Bearing displacement (BRG, mm) 25 102 – – 0.35
Superstructure unseating (UST, mm) – – 254 381 0.35

13
Bulletin of Earthquake Engineering

3 Numerical model of flared columns and bridge system

3.1 Model calibration of flared columns

Sanchez et  al. (1997) investigated the behavior of flares by conducting experiments on
forty-percent scaled bridge columns having oblong cross-sections. The oblong columns
were tested as cantilevers with flare at the bottom (upside down of actual bridge columns).
Columns with lightly reinforced flares, with flares decoupled by isolating the flare from
the superstructure, and with unflared (prismatic) sections were selected for cyclic loading
tests. In this research, four specimens tested by Sanchez et al. (1997) are a prismatic col-
umn (RDS-1), flared column (RDS-2), decoupled flare column with 25 mm thick layer of
polystyrene form between the flare and cap beam (RDS-3), and decoupled flared column
with 51  mm soffit gap (RDS-4) are selected for model calibration. These columns were
designed according to Caltrans design practice after the 1970s and have slender and lightly
reinforced flares. The geometric and reinforcing details of these specimens are shown in
Fig. 3. The cross-section of the oblong section is 640 by 914 mm and is flared into 610 by

Fig. 3  Reinforcing details and cross-section of four columns

13
Bulletin of Earthquake Engineering

1524 mm cross-section (Fig. 3). The columns are 3.96 m in height with 1.83 m tall para-
bolic flare. The main column reinforcement consists of 32 D19 bars (longitudinal reinforce-
ment ratio of 1.5%) and the transverse reinforcement consists of double interlocking D10
spirals at 51  mm pitch (transverse reinforcement ratio of 1.0%). The reinforcing steel in
the specimen was of grade 60 (414 MPa); the yield strength of longitudinal reinforcement
ranges from 465 to 472  MPa; and the yield strength of transverse reinforcement ranges
from 455 to 472 MPa. The concrete compressive strength of concrete at the day of testing
for the columns RDS-1, RDS-2, RDS-3, and RDS-4 are 33.2 MPa, 37.6 MPa, 33.6 MPa,
and 34.0 MPa, respectively.
The experimental results by Sanchez et  al. (1997) revealed that (1) the four columns
exhibited the ductile behavior (flexure failure), (2) the flared column without gap (RDS-2)
had higher maximum shear force capacity than other columns but exhibited a sudden drop of
shear force after reaching the maximum shear force as a result of damage to flare concrete,
and (3) the retrofit with gap did not produce a sudden drop of shear force in that the decou-
pled flared columns (RDS-3 and RDS-4) have very similar force–deformation characteristics
to the prismatic column (RDS-1) and thus the model of column flare outside the prismatic
portion can be negligible. Based on the experimental observations, the numerical model of
the four columns is built in OpenSees (McKenna 2011), as illustrated in Fig. 4. This figure
shows the numerical model of RDS-2 and RDS-4. Since the flare in the flared columns with
soffit gap is not modeled, the numerical model of RDS-1 and RDS-3 is essentially the same
as that of RDS-4. Element configurations are the same for all the columns except for different
fiber sections. Two and five fiber-type displacement-based beam-column elements are used for
the prismatic and flare region, respectively, of the flared column (a total of seven elements per
column). The sections in the flare region are assumed to be constant rectangular sections per
element. A total of five integration points are used for each element. Since the flare is lightly

Fig. 4  Assumptions on numerical model of flared columns without and with gap

13
Bulletin of Earthquake Engineering

confined, the confinement factor for the concrete outside the prismatic portion is ignored.
Thus, each fiber section has four different types of fibers: singly confined, doubly confined,
cover concrete, and longitudinal reinforcement. Concrete02 material model is employed to
model three concrete regions while Hysteretic material model is used to simulate the nonlinear
response of longitudinal reinforcing rebars. The lateral confining stress in each direction of the
rectangular column due to double interlocking spirals is calculated using the recommendation
of Leung and Burgoyne (2005) and two directional stresses are employed to compute the con-
fined strength ratio based on the model by Mander et al. (1988). The stress–strain relationship
of singly and doubly confined concrete based on the model of Mander et al. (1988) is assigned
to model the associated fibers.
The cross-section of a rectangular column with double interlocking spirals can be divided
into three regions: a doubly confined area, a singly confined area, and cover concrete. The
lateral stress on the inner boundary of the doubly confined area is higher than that on the outer
boundary of the singly confined area. This is associated with the uneven lateral stress distribu-
tion. To address this effect, Leung and Burgoyne (2005) proposed two modification factors
for the two confined regions to compute the lateral confining stress of concrete with double
interlocking spirals. The lateral confining stress of the singly and doubly confined section can
be expressed in a general form:
(2)
CT CT
fLZ = kLZ fL
where the subscript Z indicates the direction of the section (the directions X and Y indicate
the direction along the short and long side, respectively, of the rectangular section), the
superscript CT refers the type of confined concrete (sc = singly confined vs. dc = doubly
confined concrete. fL is the lateral confining stress for a circular column with a single spiral:
( )
2Asp s
fL = 1− (3)
ds s ds

where Asp is the cross-sectional area of spiral, ds is the diameter of confined concrete, s is
the spiral spacing. The modification factors ( kLZCT
 ) for the singly and doubly confined con-
crete in the two directions (X- and Y-directions) of the section were proposed as a function
of the ratio of the center-to-center distance of double interlocking spirals (csp) to the spiral
diameter (dsp). The factors were determined using a third-order least-squares fitting based
on the results of a set of finite element analyses by Leung and Burgoyne (2005).
The modification factors for the singly confined concrete are
( ) ( )2 ( )3
sc
csp csp csp
kLX = 1.3721 − 1.2103 + 1.5177 − 0.6835
dsp dsp dsp
( ) ( )2 ( )3 (4)
sc
csp csp csp
kLY = 1.3059 − 0.8711 + 0.8657 − 0.3017
dsp dsp dsp

and the modification factors for the doubly confined concrete are
( ) ( )2 ( )3
dc
csp csp csp
kLX = 1.9999 − 0.8869 + 1.5434 − 1.3374
dsp dsp dsp
( ) ( )2 ( )3 (5)
dc
csp csp csp
kLY = 1.9913 − 0.4046 + 0.2178 + 0.1615
dsp dsp dsp

13
Bulletin of Earthquake Engineering

To capture the cyclic deterioration of force–deformation curves of columns, this


research calibrates only the material model of longitudinal reinforcing rebars (Hysteretic
material model). Five cyclic response parameters of Hysteretic material model (px, py, dam-
age1, damage2, and β) are determined from the calibration of four numerical models based
on experimental data. The parameter px and py are the pinching factors for strain and stress,
respectively, during reloading. The parameters damage1, damage2 are related to damage
due to ductility and energy, respectively, and the parameter β is the power used to deter-
mine the degraded unloading stiffness based on ductility. These parameters used for the
model calibration are presented in Table 3. Only the ductility-related parameter damage1 is
slightly different (ranging from 0.006 to 0.007), but others are the same for all specimens.
Thus, different parameter damage1 will be used for the numerical model creation of the
bridges with different column configurations. Additionally, since the experimental columns
did not exhibit the shear failure, the effective shear stiffness of columns based on the elas-
ticity theory is used to include elastic shear deformations.
Figure 5 shows the comparison of experimental data and calibrated results for the four
columns. As observed, the simulated results exhibit good correlation with experimental
data on the inelastic behavior. Strength and stiffness deterioration are estimated with rea-
sonable accuracy. The difference between the maximum experimental and predicted lateral
force for all specimens is within 9%. Also, although the in-plane cyclic deterioration at
higher load level for RDS-3 (25 mm gap) is slightly more pronounced than that for RDS-4
(51 mm gap), the analytical predictions of RDS-3 and RDS-4 support the assumption of
neglect of flared concrete in flared column models with soffit gap (Fig. 5c, d). However, it
can be shown from Fig. 5a, b that limitations of the current numerical models are observed
for the prismatic column (RDS-1) and flared column (RDS-2). For RDS-1, the numerical
model overestimates lateral force during the second and third cycles at 8% drift. For RDS-
2, the analytical predictions underestimate the lateral force between a drift of 2% and 3%
where the flared concrete outside the prismatic portion experienced severe crushing and
cracks and spalling around the plastic hinge location during the experiment. Nevertheless,
in general, the current numerical modeling approach can reasonably predict the overall
hysteretic response of the columns in terms of stiffness, strength, and energy dissipation.

3.2 Numerical modeling of bridge system

The finite element platform OpenSees (McKenna 2011) is employed in this research to
create three-dimensional models of bridges with four column cases: (1) bridge with a
prismatic oblong column cross-section (BRIDGE-P, hereafter), (2) bridge with one-way
flared columns with a gap of 102 mm between the flare top and superstructure (BRIDGE-
G, hereafter), (3) bridge with one-way flared columns along the bridge transverse direc-
tion (BRIDGE-F1, hereafter), and (4) bridge with two-way flared columns along the bridge

Table 3  Cyclic response Specimen px py damage1 damage2 β


parameters of Hysteretic material
model for longitudinal rebars
RDS-1 0.3 0.8 0.0065 0.005 0
RDS-2 0.3 0.8 0.006 0.005 0
RDS-3 0.3 0.8 0.007 0.005 0
RDS-4 0.3 0.8 0.006 0.005 0

13
Bulletin of Earthquake Engineering

(a) (b)

(c) (d)

Fig. 5  Comparison of experimental and calibrated results for four columns

transverse and longitudinal direction (BRIDGE-F2, hereafter). The difference between the
numerical models for BRIDGE-P and BRIDGE-G is the inclusion of the flare mass outside
the prismatic portion. The numerical model for various bridge components is illustrated
in Fig.  1. The deck is modeled as a spine with several elastic elements as it is expected
to remain elastic during the earthquake loading. Columns are modeled using seven dis-
placement-based beam-column elements with fiber cross-sections following the procedure
mentioned before. The columns are divided into different sections along the height and five
integration points are used along each element to capture the nonlinear behavior of the col-
umn. Three types of stress–strain relationships of unconfined, singly and doubly confined
concrete, obtained from the incorporation of the model of Leung and Burgoyne (2005) to
the model of Mander et al. (1988), are assigned in the associated location of concrete fib-
ers with Concrete02 material model. Also, Hysteretic material model with cyclic response
parameters (px = 0.3, py = 0.8, damage1 = 0.006–0.007, damage2 = 0.005, and β = 0) for each
column shape is assigned in column longitudinal reinforcement fibers. Linearly elastic
translational spring is added at the center of the footing to simulate the behavior of foot-
ings. The passive earth pressure of the abutment is modeled using the hyperbolic model
suggested by Shamsabadi et al. (2010), while the structural response of the pile is simu-
lated using the material model of Mangalathu et al. (2016). It is assumed that the piles only
contribute to the transverse response of the abutments. Unlike integral abutment bridges,
seat abutment bridges have expansion joints between the deck and abutment. Their seismic
behavior results from the combined actions due to exterior shear keys, pounding between
deck and abutment, and elastomeric bearings, Shear keys are simulated following on the
experimental work of Silva et  al. (2009) to prohibit the transverse movement. The con-
tact material recommended by Muthukumar and DesRoches (2006) is used to produce the

13
Bulletin of Earthquake Engineering

pounding behavior and a bilinear model is to emulate the force–deformation response of


bearings. A detailed explanation of the numerical model of various components can be
found in the references (Mangalathu 2017; Mangalathu et  al. 2017a). Translational and
rotational masses are lumped at the deck, column, abutment, and footing nodes. Rayleigh
damping for the first and second vibration modes is employed for dynamic analyses.
Using the numerical modeling technique described above, statistically significant yet
nominally identical three-dimensional bridge models are generated by sampling across
the range of parameters (Table 1) using Latin Hypercube Sampling technique and are ran-
domly paired. Then, each probabilistic bridge model is also randomly paired with one of
ground motions to perform a set of nonlinear dynamic analyses for fragility analysis of
bridge classes.

3.3 Comparison of seismic column responses for bridge models

To examine the seismic behavior of bridges with column flare, deterministic bridge models
for four bridge classes are created in OpenSees using the mean value of modeling param-
eters in Table  1. Two horizontal components an earthquake record (with a peak ground
acceleration (PGA) of 0.383 g and 0.414 g) are applied to the deterministic models along
the bridge longitudinal and transverse axis, respectively, as plotted in Fig. 6. Figure 7 com-
pares the distribution of maximum column curvature ductility demands over the total col-
umn height about the bridge transverse axis for the four bridge models of integral and seat
abutment bridges. The column curvature ductility for each integration point is computed
for the idealized rectangular section along an element in the flared portion. Very similar
trends are observed for integral and seat abutment bridges. The curvature ductility distribu-
tion for BRIDGE-P and BRIDGE-G are almost identical, and thus the concrete mass out-
side the prismatic portion does not affect the column curvature. The plastic hinge occurs at
the top of the columns for the bridges with prismatic columns (BRIDGE-P and BRIDGE-
G); the curvature ductility highly concentrates on the top element (damage concentration)
and is very small for other elements. On the contrary, the plastic hinge moves down to the
middle of the columns (fourth element from the column top) for the bridges with flared
columns (BRIDGE-F1 and BRIDGE-F2); these bridges have relatively more distributed
column curvature ductility demand over the height than the bridges with only prismatic
column sections. The shift of plastic hinge location due to column flare is consistent with
the experimental observation by Sanchez et  al. (1997). In the case of integral abutment
bridges, the maximum column curvature ductility for BRIDGE-P, BRIDGE-F1, and
BRIDGE-F2 is 6.92, 6.31, and 5.12, respectively; at their maximum curvature ductility,

(a) (b)

Fig. 6  Ground motion in dynamic analysis for deterministic bridge models

13
Bulletin of Earthquake Engineering

(a) (b)

Fig. 7  Maximum column curvature ductility over the column height about the bridge transverse axis for
four bridges

the yield curvature for BRIDGE-P, BRIDGE-F1, and BRIDGE-F2 is 0.002894  rad/m,
0.002658 rad/m, and 0.002598 rad/m, respectively. Consistent with the experimental study
by Sanchez et al. (1997), the column curvature ductility decreases with the increase of the
flared cross-section.

3.4 Initial check of shear failure potential for selected bridge classes

The existence of concrete flare produces shorter shear span and thus might increase the
possibility of column shear failure. To examine the effect of shear response for the selected
bridge classes, this research initially checks the potential of shear failure by comparing the
column shear capacity with the associated demand from dynamic analyses. Since it is very
difficult for flared columns to determine the shear-span-to-depth ratio, especially the esti-
mation of shear span, this research does not employ existing displacement ductility-based
shear capacity prediction models (Sezen and Moehle 2004; Priestley et al. 1994; Kowalsky
and Priestley 2000) that are a function of the shear span. Extensive numerical and experi-
mental studies are needed to determine the appropriate definition of shear-span length of
flared columns, and are beyond the scope of the current study. For this reason, this research
employs an ACI equation (ACI Committee 318 2011) for columns subjected to flexure and
compression with modification to the shear resistance by transverse reinforcement [the sec-
ond term in the right side in Eq.  (6)]. The second term in the equation is related to the
shear resistance of double interlocking spirals proposed by Tanaka and Park (1993).
� �� √ � � � � Asp fyh
Nu fc 𝜋 �
Vn = Vc + VS = 1 + bw d + D + 2csp (6)
14Ag 6 2 s

where Nu is the compressive load on the column, fc̍’ is the compressive concrete strength,
Ag is the gross cross-sectional are of the column, bw is the width of the column, d is the
effective depth of the column, and D’ is the diameter of the circular array of longitudinal

13
Bulletin of Earthquake Engineering

reinforcement. Typically, bwd is assumed to be 0.8Ag. In Eq. (6), the shear resistance was
obtained under the assumption that only the peripheral spiral bars outside the area of inter-
lock are effective to carry the horizontal shear and thus the effectiveness of the peripheral
spiral bars is almost equivalent to that of the transverse reinforcement in an equivalent rec-
tangular section. Figure 8 shows the normalized transverse column shear force demand to
the associated capacity (Eq. 6) for integral and seat abutment bridges with one-way column
flare. The column shear force demand is the maximum column shear force of the columns
obtained from 320 dynamic analyses. For a conservative estimate, the prismatic section
is selected to compute the shear force capacity. It is seen from the figures that the column
shear force demand did not reach the column shear capacity for 320 bridge-ground motion
pairs (maximum of normalized shear force demands is within 0.7). For this reason, this
research does not account for the shear response of bridge columns. Instead, the elastic
shear stiffness is included in inelastic beam-column elements. However, the shear demand-
to-capacity ratio less than one is limited for the selected bridge classes. Other bridges
classes not considered in this research might be susceptible to the shear failure of flared
columns, and further studies are needed to address this issue.

4 Fragility results for bridge models

Fragility curves are generated through the methodology presented before for four column
classes (BRIDGE-P, BRIDGE-G, BRIDGE-F1, and BRIDGE-F2) of integral and seat
abutment bridges. The fragility curves for bridge components of integral and seat abut-
ment bridges are plotted in Figs. 9 and 10, respectively. Since the shift in the component
fragility curves for integral abutment bridges associated with column flare is a very similar
trend to those for seat abutment bridges, the fragility curves for integral abutment bridges
are not shown here, but their fragility characteristics are summarized in Tables  4 and 5.
These tables present the median and dispersion value, respectively, of fragilities for various
bridge components and bridge system. It is noted from Figs. 9 and 10 and Tables 4 and 5
that the column flare makes the column less vulnerable and the influence of flare is more
significant at higher limit states. Following inferences can be obtained from Figs. 9 and 10
and Tables 4 and 5.

• There is little or no variation between the fragilities for the prismatic case (BRIDGE-
P), and column flare with a gap case (BRIDGE-G), for all the limit states. The case is

(a) (b)

Fig. 8  Comparison of transverse column shear force demands and capacities for bridges with one-way col-
umn flare

13
Bulletin of Earthquake Engineering

(a) (b)

(c) (d)

(e)

Fig. 9  Comparison of component fragility curves for integral abutment bridges: a COL (lower limit state),
b COL (upper limit state), c ABP, d ABA, e ABT

true irrespective of the type of abutments: integral or seat abutment bridge. It can thus
be concluded that the decoupled flared column essentially behaves the same as pris-
matic column. This observation is consistent with the experimental results by Sanchez
et al. (1997). Therefore, the decoupling of flare by sawing away the steel and concrete
at the soffit is a retrofit solution for columns susceptible to shear failure due to flare,
although the shear failure is not addressed in this study.
• The prismatic column is more vulnerable than the flared columns for the selected
bridge cases. This observation is contrary to the popular belief that the flare always
makes the column vulnerable. The less vulnerability of the flared columns considered
in this research is attributed to the increased flexural capacity of the columns due to the
flare and less susceptibility of the selected bridge columns to shear failure. Note that
this conclusion is valid only for the selected bridge classes.
• The column flare reduces the transverse response of bridge components such as the
column (COL), transverse abutment action (ABT), and bearing (BRG). It is attributed
to the change in stiffness and the flexural capacity of the columns with the flare in the
transverse direction. However, the flare has little effect on the longitudinal response of

13
Bulletin of Earthquake Engineering

(a) (b)

(c) (d)

(e) (f)

(g)

Fig. 10  Comparison of component fragility curves for seat abutment bridges: a COL (lower limit state), b
COL (upper limit state), c ABP, d ABA, e ABT, f UST, g BRG

bridge components such as the passive and active abutment actions and unseating. It is
due to the higher stiffness associated with the passive abutment action and pounding
effect.
• The seismic fragility of all the seat-abutment bridges is governed by bearings at the
slight and moderate damage states (­LS1 and ­LS2). On the other hand, for all integral
abutment bridges, the transverse abutment action governs the bridge system response

13

13
Table 4  Median value of system and component fragility curves (median ­Sa−1s in g)
LS LS1 LS2 LS3 LS4
Bridge Pa G F1 F2 P G F1 F2 P G F1 F2 P G F1 F2

Integral abutment bridges


 SYS 0.188 0.184 0.221 0.220 0.596 0.594 0.742 0.747 1.067 1.073 1.549 1.548 1.311 1.314 1.922 1.920
 COL 0.281 0.281 0.379 0.379 0.791 0.791 1.124 1.122 1.070 1.071 1.544 1.541 1.314 1.315 1.915 1.911
 ABP 1.350 1.354 1.403 1.452 4.800 4.845 5.025 5.211 – – – – – – – –
 ABA 0.588 0.588 0.609 0.630 1.725 1.733 1.799 1.866 – – – – – – – –
 ABT 0.194 0.194 0.230 0.229 0.661 0.660 0.805 0.803 – – – – – – – –
Seat abutment bridges
 SYS 0.108 0.107 0.128 0.134 0.434 0.430 0.550 0.555 0.847 0.839 1.213 1.175 1.054 1.048 1.521 1.472
 COL 0.199 0.199 0.279 0.276 0.610 0.608 0.869 0.850 0.845 0.842 1.212 1.180 1.054 1.050 1.518 1.474
 ABP 1.143 1.136 1.168 1.239 3.107 3.075 3.088 3.276 – – – – – – – –
 ABA 0.889 0.872 0.901 0.924 2.794 2.700 2.772 2.824 – – – – – – – –
 ABT 0.216 0.215 0.254 0.257 0.636 0.633 0.761 0.767 – – – – – – – –
 UST – – – – – – – – 7.794 7.663 7.177 7.280 15.687 15.356 13.987 13.938
 BRG 0.111 0.111 0.130 0.137 0.521 0.519 0.667 0.689 – – – – – – – –
a
 P, G, F1, and F2 indicate BRIDGE-P, BRIDGE-G, BRIDGE-F1, and BRIDGE-2F, respectively
Bulletin of Earthquake Engineering
Bulletin of Earthquake Engineering

Table 5  Dispersion of system Bridge Integral abutment bridge Seat abutment bridge


and component fragility curves
Pa G F1 F2 P G F1 F2

SYSb 0.544 0.546 0.618 0.617 0.536 0.534 0.606 0.622


COL 0.550 0.551 0.624 0.634 0.536 0.537 0.594 0.631
ABP 0.931 0.931 0.936 0.925 0.649 0.644 0.649 0.680
ABA 0.909 0.909 0.916 0.907 0.815 0.801 0.797 0.806
ABT 0.545 0.543 0.606 0.604 0.574 0.570 0.619 0.621
UST – – – – 0.903 0.898 0.867 0.851
BRG – – – – 0.581 0.575 0.654 0.651
a
 P, G, F1, and F2 indicate BRIDGE-P, BRIDGE-G, BRIDGE-F1, and
BRIDGE-2F, respectively
b
 The dispersion value for the system is the average of the dispersions
for ­LS1, ­LS2, ­LS3, and ­LS4

at these damage states. For the extensive and complete damage states ­(LS3 and ­LS4),
the column fragility dominates the system fragility of both integral and seat abutment
bridges. These observations are consistent with the previous studies on prismatic bridge
columns in California (Mangalathu 2017; Mangalathu et al. 2018b).
• There is not much statistical variation between the median value of fragilities for
bridges with one-way flare (BRIDGE-F1) and two-way flare (BRIDGE-F2). The
increase of the stiffness associated with column flare does not influence the longitudinal
response of bridge components due to the high resistance of abutment backfill.

5 Summary and conclusions

The increased aesthetic appeal of flared columns accelerated the construction of bridge
columns with flares after the 1970s in California. Although columns with flares provide
flexural over-strength at the top of columns, it has been traditionally thought that the shift
of plastic hinge location makes the flared column more vulnerable. This paper investigates
the seismic vulnerability of three-span bridges with and without flares having oblong (rec-
tangular) cross-sections in California through the development of fragility curves. This is
the first attempt to examine the seismic response of bridges with flared oblong column
cross-sections.
The oblong column model with and without flares is firstly calibrated with existing
experimental data available in the literature. Three different configurations such as pris-
matic, flared column without a gap, and flared column with a gap (four specimens) are used
for model calibration. The analytical predictions realistically capture the overall response
of the columns with regard to the stiffness, strength, and energy dissipation. The calibrated
column model is integrated into the bridge system and probabilistic bridge system models
are created for bridges with flared and non-flared three-span, two-column bent configura-
tions by reflecting the material, geometric and structural uncertainties in California. Non-
linear dynamic analyses are then performed for a set of the bridges models with a suite of
ground motions. The seismic demands obtained from the dynamic analyses are used to
develop demand models as function of the ground intensity measure. The demand models
are convolved with the capacity models to derive fragility curves for the selected bridge

13
Bulletin of Earthquake Engineering

classes. The fragility curves are generated for four different bridge column configurations:
(1) bridge with a prismatic oblong column cross-section, (2) bridge with one-way flared
columns with a gap of 102 mm between the flare top and superstructure, (3) bridge with
one-way flared columns along the bridge transverse direction, and (4) bridge with two-
way flared columns along the bridge transverse and longitudinal direction. It is noted that
the bridges with flare gap are essentially behaves the same as prismatic columns and the
decoupling of the flare is a retrofitting solution to change the behavior of flared columns
to prismatic columns. Additionally, if the column is not susceptible to shear failure, the
seismic vulnerability of the flared columns is less than the prismatic columns and thus the
flare makes a beneficial contribution to the mitigation of seismic damage by increasing the
stiffness and strength of the bridge system. For the selected bridge classes, there is little or
no statistical difference in the median value of fragilities for columns with one-way and
two-way flares due to the large resistance of abutment backfill. Note that the observations
in this research are based on the selected bridge configurations (straight three-span bridges
with oblong cross-sections and two-column bents). For the irregular bridges with skewness
and horizontal curvature, the flare might be detrimental, and further studies are needed to
evaluate the effect of skewness and horizontal curvature on flared columns. Further studies
are needed to investigate the effect of flare on other bridge configurations, especially with
the potential of column shear failure.

Acknowledgements The research was supported by a Grant (18CTAP-C130227-02) from Technology


Advancement Research Program (TARP) funded by Ministry of Land, Infrastructure and Transport of
Korean government. This work is also financially supported by Ministry of Public Administration and Secu-
rity as Disaster Prevention Safety Human Resource Development Project.

References
ACI Committee 318 (2011) Building code requirements for reinforced concrete (ACI 318-11) and commen-
tary. American Concrete Institute, Farmington Hills
Baker JW, Lin T, Shahi SK, Jayaram N (2011) New ground motion selection procedures and selected
motions for the PEER transportation research program. PEER report 2011/03, Pacific Earthquake
Engineering Research Center, University of California, Berkeley, USA
Banerjee S, Shinozuka M (2008) Mechanistic quantification of RC bridge damage states under earthquake
through fragility analysis. Probab Eng Mech 23(1):12–22
Caltrans (2006) Seismic design criteria version 1.4. Office of Structures Design, California Department of
Transportation, Sacramento
Cornell C, Jalayer F, Hamburger R, Foutch D (2002) Probabilistic basis for 2000 SAC Federal Emergency
Management Agency steel moment frame guidelines. J Struct Eng 128(4):526–533
Correal JF, Saiid S, Sanders D, El-Azazy S (2007) Shake table studies of bridge columns with double inter-
locking spirals. ACI Struct J 104(4):393–401
Dimitrakopoulos EG, Paraskeva TS (2015) Dimensionless fragility curves for rocking response to near-fault
excitations. Earthq Eng Struct Dyn 44(12):2015–2033
Huo YL, Zhang J (2013) Effects of pounding and skewness on seismic responses of typical multispan high-
way bridges using the fragility function method. J Bridge Eng 18(6):499–515
Jeon J-S, Mangalathu S, Song J, DesRoches R (2017) Parameterized seismic fragility curves for curved
multi-frame concrete box-girder bridges using Bayesian parameter estimation. J Earthq Eng. https​://
doi.org/10.1080/13632​469.2017.13422​91
Kowalsky MJ, Priestley MJN (2000) Improved analytical model for shear strength of circular reinforced
concrete columns in seismic regions. ACI Struct J 97(3):388–396
Leung HY, Burgoyne CJ (2005) Uniaxial stress-strain relationship of spirally confined concrete. ACI Mater
J 102(6):445–453
Mackie K, Stojadinovic B (2001) Probabilistic seismic demand model for California highway bridges. J
Bridge Eng 6(6):468–481

13
Bulletin of Earthquake Engineering

Mander JB, Priestley MJN, Park R (1988) Theoretical stress–strain model for confined concrete. J Struct
Eng 114(8):1804–1826
Mangalathu S (2017) Performance based grouping and fragility analysis of box-girder bridges in California.
PhD dissertation, Georgia Institute of Technology, USA
Mangalathu S, Jeon J-S, DesRoches R, Padgett JE (2016) ANCOVA-based grouping of bridge classes for
seismic fragility assessment. Eng Struct 123:379–394
Mangalathu S, Jeon J-S, Padgett JE, DesRoches R (2017a) Performance-based grouping methods of bridge
classes for regional seismic risk assessment: application of ANOVA, ANCOVA, and non-parametric
approaches. Earthq Eng Struct Dyn 46(14):2587–2602
Mangalathu S, Soleimani F, Jeon J-S (2017b) Bridge classes for regional seismic risk assessment: improv-
ing HAZUS models. Eng Struct 148:755–766
Mangalathu S, Heo G, Jeon J-S (2018a) Artificial neural network based multi-dimensional fragility develop-
ment of skewed concrete bridge classes. Eng Struct 162:166–176
Mangalathu S, Jeon J-S, DesRoches R (2018b) Critical uncertainty parameters influencing seismic perfor-
mance of bridges using Lasso regression. Earthq Eng Struct Dyn 47(3):784–801
McKenna F (2011) OpenSees: a framework for earthquake engineering simulation. Comput Sci Eng
13(4):58–66
Monteiro R, Delgado R, Pinho R (2016) Probabilistic seismic assessment of RC bridges: part I—uncertainty
models. Structures 5:258–273
Muthukumar S, DesRoches R (2006) A Hertz contact model with non-linear damping for pounding simula-
tion. Earthq Eng Struct Dyn 35(7):811–828
Nada HM, Sanders DH, Saiidi MS (2003) Seismic performance of RC bridge frames with architectural-
flared columns. CCEER 03-03, Center for Earthquake Engineering Research, University of Nevada,
USA
Nielson BG, DesRoches R (2007) Seismic fragility methodology for highway bridges using a component
level approach. Earthq Eng Struct Dyn 36(6):823–839
Pan Y, Agrawal AK, Ghosn M (2007) Seismic fragility of continuous steel highway bridges in New York
State. J Bridge Eng 12(6):689–699
Priestley MJN, Verma R, Xiao Y (1994) Seismic shear strength of reinforced concrete columns. J Struct
Eng 120(8):2310–2329
Priestley MJN, Seible F, Calvi G (1996) Seismic design and retrofit of bridges. Wiley, New York
Ramanathan KN (2012) Next generation seismic fragility curves for California bridges incorporating the
evolution in seismic design philosophy. PhD dissertation, Georgia Institute of Technology, USA
Rogers LP, Seo J (2017) Vulnerability sensitivity of curved precast-concrete I-girder bridges with various
configurations subjected to multiple ground motions. J Bridge Eng 22(2):04016118
Saiidi MS, Wehbe NI, Sanders DH, Caywood CJ (2001) Shear retrofit of flared RC bridge columns sub-
jected to earthquakes. J Bridge Eng 6(3):189–197
Sanchez AV, Seible F, Priestley MJN (1997) Seismic performance of flared bridge columns. SSRP-97/06,
Structural Systems Research Project, University of California, San Diego, USA
Seo J, Linzell DG (2012) Horizontally curved steel bridge seismic vulnerability assessment. Eng Struct
34:21–32
Seo J, Rogers LP (2017) Comparison of curved prestressed concrete bridge population response between
area and spine modeling approaches toward efficient seismic vulnerability analysis. Eng Struct
150:176–189
Sezen H, Moehle JP (2004) Shear strength model for lightly reinforced concrete columns. J Struct Eng
130(11):1692–1703
Shamsabadi A, Khalili-Tehrani P, Stewart JP, Taciroglu E (2010) Validated simulation models for lateral
response of bridge abutments with typical backfills. J Bridge Eng 15(3):302–311
Silva PF, Megally S, Seible F (2009) Seismic performance of sacrificial exterior shear keys in bridge abut-
ments. Earthq Spectra 25(3):643–664
Soleimani F, Mangalathu S, DesRoches R (2017) A comparative analytical study on the fragility assessment
of box-girder bridges with various column shapes. Eng Struct 153:460–478
Tanaka H, Park R (1993) Seismic design and behavior of reinforced concrete columns with interlocking
spirals. ACI Struct J 90(2):192–203

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

13

You might also like