You are on page 1of 16

Engineering Structures 184 (2019) 99–114

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Collapse capacity of reinforced concrete skewed bridges retrofitted with T


buckling-restrained braces
Yuandong Wanga,b, Luis Ibarraa, , Chris Pantelidesa

a
Department of Civil and Environmental Engineering, University of Utah, Salt Lake City, UT 84112, USA
b
BHB Consulting Engineers, Salt Lake City, UT 84115, USA

ARTICLE INFO ABSTRACT

Keywords: This study assesses the collapse capacity and failure modes of skewed bridges retrofitted with buckling-re-
Collapse capacity strained braces (BRBs) at the column bent. The asymmetric configuration of these bridges requires a three-
Three-dimensional dimensional numerical model that considers uncertainties in the seismic ground motions. The factors controlling
Incremental dynamic analysis the seismic performance of these bridges are obtained from a case study of a three-span reinforced concrete box
Skewed bridge
girder skewed bridge with varying skew angles of 0°, 18°, 36°, and 54°. Numerical models with distributed
Retrofit
Buckling-restrained brace
plasticity and concentrated plasticity are created considering material deterioration and plastic deformation
properties. A numerical distributed plasticity model with strength and stiffness deterioration is calibrated with a
concentrated plasticity model by performing two-dimensional incremental dynamic analyses (IDAs) on the
straight bridges (i.e., 0° skew angle), using 21 far-field ground motions. Then, the collapse capacity of original
and retrofitted skewed bridges is obtained from three-dimensional IDAs using the distributed plasticity model.
The collapse capacity and failure modes of BRB components are summarized based on investigations from ex-
perimental data. Nonlinear time history analyses indicate that the BRB retrofit greatly improves the seismic
performance of skewed bridges, but it has negligible effects on the bridge’s collapse capacity after BRB failure.
However, the use of BRB greatly reduces the bridge’s probability of failure, and the mean annual frequency of
global collapse, since the BRB components dissipate seismic energy as structural fuses. The proposed method of
parameter calibration, including BRB failure, is found to be sufficiently reliable to perform satisfactory results.

1. Introduction (RTR) variability, which accounts for the different frequency content of
the applied records. The IDA method is commonly applied to two-di-
The performance of skewed bridges retrofitted with buckling-re- mensional (2D) numerical models, to include the 2D hysteretic non-
strained braces (BRBs) was evaluated by Wang et al. [1,2] under linear spring models, which incorporate strength and stiffness dete-
seismic records with accelerations matching the maximum considered rioration [3] that is able to detect structural collapse. For instance,
earthquake (MCE) in Ripon, CA, USA. The study showed that BRB Altoontash [6], Haselton [7], Lignos et al. [8], and Lignos and Kra-
components improve the seismic performance of bridges under service winkler [9] implemented and improved the deterioration model with
and strength limit states by decreasing drifts in the bents, and by re- peak-oriented hysteretic responses [3] to represent the behavior of RC
ducing the steel and concrete strains of the original reinforced concrete beams and columns in OpenSees [10]. The accuracy of these models can
(RC) columns. The present study evaluates the global collapse capacity be attributed to calibration of hundreds of experiments from 2D bidir-
of the retrofitted bridges. Global collapse indicates that the structural ectional steel and concrete component tests. Even though the 2D model
system is unable to maintain its gravity loading-carry capacity in the with concentrated plasticity can represent the main characteristics of
presence of lateral loading, as obtained from incremental dynamic regular structures, three-dimensional (3D) numerical models are still
analyses (IDAs) [3–5]. needed to analyze asymmetrical structures, such as skewed or curved
In the IDA method, nonlinear time history analyses (THAs) of the bridges, as well as structures that require 3D structural interaction and
structure are performed under monotonically increased intensity levels biaxial seismic loading.
of the applied ground motion. A set of seismic records is typically used Bidirectional concentrated plasticity spring model and distributed
to obtain the variance of collapse capacity due to Record-To-Record plasticity fiber model have been used to simulate seismic structural


Corresponding author.
E-mail addresses: matt.wang@utah.edu (Y. Wang), luis.ibarra@utah.edu (L. Ibarra), c.pantelides@utah.edu (C. Pantelides).

https://doi.org/10.1016/j.engstruct.2019.01.033
Received 14 January 2018; Received in revised form 16 December 2018; Accepted 8 January 2019
Available online 23 January 2019
0141-0296/ © 2019 Elsevier Ltd. All rights reserved.
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 1. Concentrated plasticity and distributed plasticity (adapted from Deierlein et al. [36]).

response of irregular structures. For example, Vamvatsikos and Cornell load [29–33]. In this study, however, bridge collapse failure in related
[12] utilized two nonlinear springs along the two major horizontal to a sidesway collapse failure mode [34], in which the bridge loses its
structural axes to evaluate the seismic performance of a 20-story steel ability to withstand gravity loads under the presence of seismic loading,
frame in 3D realization. The 3D response parameters were obtained and the lateral displacements grow unbounded, flattening the IDA
from a post-processing approach that considers the square root of the curves. In the process of the bridge reaching this global structural
sum of squares (SRSS) of the maximum peak drift from each individual collapse, different types of material and component failures may take
direction. Fathieh [13] performed IDAs of modular steel buildings, place, such as concrete crushing, rebar tensile rupture, and rebar
utilizing a distributed plasticity fiber element model with plastic hinges buckling, among others.
to simulate the nonlinear behavior of beams and columns. In a bridge
engineering application, and due to the irregular configuration of the 2. Methodology to evaluate bridge collapse capacity
investigated case bridge, Vamvatsikos and Sigalas [14] generated a 3D
curved bridge model with distributed plasticity fiber elements, and then Seismic performance assessment requires computation of para-
performed IDAs. In a similar way, Billah and Alam [15], and He et al. meters, such as strength, drift, and ductility, which should meet per-
[16] carried out IDAs to evaluate the seismic performance of retrofitted formance targets and correlate to structural damage [35]. Concentrated
RC bridges using distributed plasticity models under bidirectional plasticity models utilize 2D nonlinear springs to represent the inelastic
ground motions. However, concentrated plasticity and distributed deformations at the end of each element, whereas the rest of the
plasticity models have not been previously compared to evaluate their component only exhibits elastic deformations. In the concentrated
capability to predict collapse capacity. plasticity model, the concentrated plastic zero-length hinge that in-
Previous numerical and experimental studies showed that BRB cludes the moment-rotational model parameters works in series with
components in the bridge bent assist the structure in dissipating seismic the elastic components. The concentrated plasticity model is numeri-
energy, thus improving the seismic performance of bridge bents in the cally efficient, and less prone to error, but it cannot represent cyclic
transverse direction [2,17–22]. BRBs have also been proposed to in- deterioration under bidirectional seismic loading.
crease the seismic capacity of bridge decks in the longitudinal direction The distributed plasticity models in OpenSees [10] include a finite
of bridge [23–27]. However, few studies have assessed the ultimate length hinge model and a fiber model [35,36], as shown in Fig. 1. The
capacity of BRBs, especially under seismic events with high magnitude. distributed plasticity elastic element with a finite length hinge model
Andrews et al. [28] developed an analytical method to predict the cu- consolidates the nonlinear moment-curvature relationship in the in-
mulative plastic ductility (CPD) capacity of BRB components for steel elastic hinge region, producing overall spread integration of deforma-
frames, utilizing a statistical framework and past experimental results tions along the hinge. The fiber element model simulates the properties
from 76 quasi-static BRB tests. of the RC element by distributing plasticity into several numerical in-
In this study, IDAs are carried out to investigate the collapse capa- tegration sections (five in this study) that include material fibers, to
city of RC skewed bridges before and after being retrofitted with BRBs. represent the uniaxial stress-strain characteristics of reinforcing steel
For this purpose, the material properties of distributed plasticity fiber and concrete. Each fiber section is assumed to remain plane while the
models are calibrated with experimental results that include material individual fibers are numerically integrated to obtain the stress, mo-
deterioration. In this study, material deterioration refers to strength and ment curvature, and axial force-strain relations.
stiffness deterioration of the component after reaching a peak strength.
A 3D distributed plasticity model is first calibrated with a 2D con- 2.1. Prototype bridge
centrated plasticity model in isolated RC columns and three-column
bridge bents, which are subjected to monotonic and quasi-static loading The evaluated system is a three-span cast-in-place RC box girder
protocols, as well as THAs. The influence of skew angle and BRB retrofit skewed bridge with three columns in each bent. The model is modified
is then studied by performing 3D IDAs with the validated distributed from an existing bridge with a total length of 127.5 m and a skew angle
plasticity fiber models. A methodology is also presented to define the of α = 36°, located in Ripon, California, USA [11,37]. In the transverse
material and analysis parameters for the 3D modeling. The benefit of direction, the bridge bent has a length of 23.347 m, and three circular
BRB retrofit is further investigated by comparing the annual frequency columns with a diameter of 1.68 m, and a height of 7.38 m, as shown in
of collapse of the original bridges and the BRB-retrofitted bridges in 3D Fig. 2. The columns have a design concrete compressive strength
realizations. f 'c = 34.5 MPa, longitudinal reinforcement arranged in bundles of two
Traditionally, bridge column failure in design practice is associated rebars with a total of 34 No. 14 rebars (43 mm in diameter), and No. 6
to fracture of column bars, core concrete crushing failure, or a 15% (19 mm in diameter) spiral reinforcement with an 85 mm spacing. The
reduction in the lateral load carrying capacity with respect to the peak columns are pin-connected to the foundation with a rebar hinge [38],

100
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 2. Column bent design details (adapted from Caltrans [37]).

and only the interior longitudinal reinforcement is continuously con- deteriorating parameters are obtained from calibration of moment-ro-
nected to the foundation, with an expansion joint filler injected into the tations and hysteretic curves with member experimental tests in uni-
column edges (Fig. 2). The piers are supported on 24H-shaped directional loading protocols, with limitation of 2D application. The
305 × 79 steel bearing piles (pile width = 305 mm, pile model includes a backbone curve with a negative post-capping stiffness
weight = 79 kg/m) per column, and the seat-type abutments have 9 associated with reinforcement buckling and fracture, concrete crushing,
bearing pads and 40 piles underneath. and bond failure. Four types of cyclic deterioration can be simulated:
The bridge has a seat-type abutment with wingwalls at both sides of basic strength deterioration, post-capping strength deterioration, un-
the abutment stemwall, and pile foundations underneath. To account loading stiffness deterioration, and accelerated reloading stiffness de-
for the effects of the soil filling in between the wingwall and the terioration. The empirical parametric equations used to define the RC
stemwall, nonlinear springs are included in the model in the long- deterioration model were developed by Haselton and Deierlein [42],
itudinal and skew direction. The transverse stiffness of the embankment assuming a lognormal distribution for the evaluated parameters.
and the wedge are estimated by multiplying the dynamic stiffness of the For columns, the ratio of the equivalent elastic stiffness, EIy , to the
embankment by an effective length [39], with the embankment soil fill gross flexural stiffness, EIg ; is defined by Eq. (1):
stiffness, as recommended by Caltrans [40]. The shear keys at the
abutment are designed according to AASHTO [41] and Caltrans [40]. EIy P
= 0.065 + 1.05
There are elastic bearings between the bridge’s deck and abutment that EIg A g f 'c (1)
provide flexibility in the vertical direction. The bridge is expected to
meet life safety requirements, but operational limits may be exceeded where P is the axial load, A g is the cross-section area, is the concrete
f 'c
P
under the design basis earthquake. Thus, seismic retrofit may be needed compressive strength, and the axial load index, v = ' . The input
A gf c
to keep the bridge seismic performance under operational limits, if parameters of a concentrated plasticity model are shown in Fig. 3,
bridge downtimes were deemed unacceptable, as in the case of bridges where the yield rotation and ultimate rotation are defined by Eqs. (2)
providing access to hospitals or other essential facilities. and (3), respectively:
tot p
y = cap cap (2)
2.2. Deterioration models with concentrated plasticity
tot
c = cap (3)
The concentrated plastic model is composed of an elastic element
that represent the physical length of structural member, and a spring Eq. (4) estimates the plastic rotation capacity ( cap ),
p
based on sta-
that captures the component nonlinear degradation based on the de- tistically significant variables, including bond-slip, a sl , as defined by
teriorating hysteretic modes proposed by Ibarra and Krawinkler [34]. Fardis and Biskinis [43]; the lateral confinement ratio, sh ; the rebar
Overall model stiffness equals to the actual member stiffness. The buckling coefficient, sn ; and the longitudinal reinforcement ratio, .

101
Y. Wang et al. Engineering Structures 184 (2019) 99–114

time history loading protocols. The material selection is based upon


feasibility of defining material deteriorations, and computational sta-
bility under quasi-static and extreme seismic loadings.
The RC elements in the distributed plasticity models of this study
are defined by the Concrete02 material for concrete, and hysteretic
material for steel in OpenSees [10]. The properties of the ultimate
strength for the confined concrete models are characterized with the
Mander’s model [44,45]. The strength and stiffness degradation, as well
as residual strength are based on the Modified Kent and Park method
[46]. The Mander’s model was used for the confined concrete modeling
because it is widely used, but it does not have some of the features
needed to accurately reproduce the behavior of unconfined concrete.
The compression properties of the unconfined concrete are defined by
the Mander model [44,45], as shown in Fig. 4a. As observed in ex-
periments [47–50], the steel degradation in this study includes the
softening, buckling after peak strain, and tensile rupture of rebars, as
Fig. 3. Monotonic curve of the concentrated plasticity model and distributed
presented in Fig. 4b and Table 2. The buckling properties are based on
plasticity model.
the parameters defined by Dhakal and Maekawa [49]. Table 2 sum-
marizes the parameters for the distributed plasticity fiber model, ac-
'
p
cap = 0.12(1 + 0.55a sl)0.16v (0.02 + 40 sh )
0.430.540.01cunits f c (0.66)0.1s n cording to the design parameters of the investigated bridge. The bond
slip contribution at the top of the columns are not included in this fiber
(2.27)10.0 (4)
model [50], considering the contribution of bond-slip on a well-de-
The total rotation capacity ( cap ),
defined by Eq. (5), refers to the
tot
signed cast-in-place bridge specimen is not significant [51]. Further-
rotation at the capping (peak) point considering elastic and plastic more, bond slip may become more relevant when the longitudinal bars
deformations. are damaged [52], but then above mentioned material deteriorations
' should dominate the concrete and steel damages in the plastic hinge.
tot
= 0.12(1 + 0.4a sl)0.2v (0.02 + 40 0.520.560.01cunits f c (2.37)10.0
cap sh ) (5) Also, the BRB retrofit effects should not be affected by bond slip during
The post-capping rotation capacity ( IDA analysis.
pc ) is defined as,

pc = 0.76(0.031)v (0.02 + 40 sh )1.02 0.10 (6) 2.4. Component model comparison


whereas the hardening stiffness that is defined as the ratio of
To validate the 3D deterioration model, results from the distributed
maximum moment capacity to yield moment capacity, Mc /My , is ob-
plasticity model are compared to those obtained from the 2D con-
tained from Eq. (7).
centrated plasticity models of both individual columns and three-
column bents, under unidirectional monotonic and cyclic loading. The
'
Mc /My = 1.25(0.89)v 0.910.01cunitsf c (7)
cap beam between columns was modeled using linear-elastic beam-
Also, c units is a unit conversion variable equal to 1.0 when is in the f 'c column elements, and its mass and moment of inertia were calculated
SI units (MPa). Eq. (8) denotes the cyclic and stiffness deterioration and based on its net area. Also, the column segment inserted into the cap
the energy dissipation capacity of a RC column element. Table 1 sum- beam was assumed to be a rigid link. The hysteretic curves of the
marizes the parameters for the concentrated plasticity model, according concentrated and distributed plasticity models are shown in Figs. 5a
to the design parameters of the investigated bridge. and 5b for an individual stand-alone column with fixed base, and for a
= 170.7(0.27)v (0.10) s/d (8) three-column bridge bent with base pin connections, respectively. The
hysteretic curves from these numerical models are close, and predict
similar hysteretic properties, such as elastic stiffness, post-capping
2.3. Deterioration models with distributed plasticity stiffness, and cyclic deterioration properties. The cyclic responses of the
columns with concentrated and distributed plasticity models under the
The distributed plasticity model includes several numerical in- loading protocols in Nguyen et al. [53] are shown in Fig. 5a. In both
tegration sections that are composed of several fiber sections with models, the column strength loss results in similar maximum dis-
certain material properties. Its model structure could be used in 3D IDA placements at peak strength and negative post-capping stiffness ratio,
with proper deteriorations defined in the material level. In the pre- which are two important parameters for predicting structural in-
liminary stage of this study, all available steel and concrete material stability. A plain load pattern with a linear time series is used for
elements in OpenSees [10] were experimentally utilized in a fix-base pushover analysis with the lateral load application in the transverse
RC column and three-column RC bents under monotonic and cyclic, and direction of bent.
The quasi-static loading protocols applied to individual columns are
Table 1 also used on an individual three-column bridge bent, as shown in
Summary of the parameters for the concentrated Fig. 5b. The strength of the distributed model shows abrupt deteriora-
plasticity model.
tion at some cycles due to tensile rupture of the column’s longitudinal
Parameter Value reinforcement, while the concentrated plasticity model has a constant
post-capping stiffness. Even though the two models predict a similar
EIy 0.2
EIg
(Eq. (1)) onset of structural instability under quasi-static loading, results from
p
cap (Eq. (2)) 0.043 THAs are also compared to ensure their accuracy.
0.063 Note that concentrated plasticity models can only be defined in one
cap (Eq. (3))
tot

0.25 direction. Thus, a comparison of concentrated and distributed plasticity


pc (Eq. (4))
Mc /My (Eq. (5)) 1.16 models is only possible for straight bridges. For this reason, in the case
(Eq. (6)) 140 of skew bridges the comparison is only performed for the concrete bents
in the transverse bridge direction. Figs. 6a and 6b present the cyclic

102
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 4. Stress-strain curves for the distributed plasticity model: (a) Concrete; (b) steel.

Table 2
Summary of the parameters for the distributed plasticity model.
Confined Concrete Parameter Unconfined Concrete Parameter Steel Parameter

Compressive strength 44.9 MPa Compressive strength 34.5 MPa Yield strength 470 MPa
Strain at Max. strength −0.0028 Strain at Max. strength −0.0036 Ultimate strength 655 MPa
Crushing strength 8.3 MPa Crushing strength 0 MPa Elastic modulus 2 × 105 MPa
Strain at crushing strength −0.0550 Strain at crushing strength −0.0050 Strain hardening ratio 0.01
Strain at Max. strength 0.060
Max. tensile strain [47,48] 0.15
Intermediate stress [49] 510 MPa
Intermediate strain [49] 0.106
Max. buckling strain [49] 0.146

drift-bent force obtained from the concentrated and distributed plasti- follow-up study. The records represent 14 events, eight were California
city models, when subjected to ground motions from Loma Prieta and earthquakes and 6 were from five different foreign countries, which
Imperial Valley earthquakes. The ground motions were scaled to include 16 records classified as Site Class D (stiff soil sites), in ac-
seismic intensities of SaSRSS = 3.2 g and 6.0 g. As observed, the hys- cordance with the bridge’s soil conditions.
teretic responses are close with differences smaller than 5%, even when For each ground motion set, both horizontal records are first rotated
considering strength and stiffness deterioration, verifying the accuracy to the principal directions that provides the minimum correlation
of the distributed plasticity model to simulate component strength de- coefficient of rotated accelerations when each pair of horizontal ac-
terioration. celeration time histories are rotated counterclockwise by two dimen-
sional rotation matrix at each second. For instance, the original accel-
erations ax (t) and ax (t) , are transformed into the rotated accelerations
2.5. Ground motion selection and scaling a 'x (t) and a 'y (t) , as follows:

The bridge numerical analyses included 21 far-field records selected a 'x (t) = a x (t)cos + a y (t)sin (9a)
from FEMA P695 [54], with a mean distance of 16.4 km, which meets
the site condition of case bridge. Also, the BRB retrofit of skewed
bridges under near-field ground motions will be investigated in a a 'y (t) = a x (t)sin + a y (t)cos (9b)

Fig. 5. Cyclic behavior of component models used in this study: (a) column; (b) bent.

103
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 6. Bridge bent time history responses under different GMs and diferernt intensities: (a) Loma Prieta Sa SRSS = 3.2 g; (b) Imperial_Valley Sa SRSS = 6.0 g.

where is the rotation angle. In the resulting principal directions, the direction to keep the analysis in the plane, and be able to apply the
ratio cov(x ', y ')
is approximately zero, where cov(x ', y ') is the covariance of concentrated plasticity model. Structural global P-Δ effects [56] are
x'× ' y considered in these analyses by applying P-Δ coordinate transforma-
rotated acceleration, and x' , y ' are the standard deviations of the ro- tions of each bridge column. Note that element P-δ effects are ne-
tated horizontal accelerations. glected.
The major principal direction represents the record with the larger The original bridge model with distributed plasticity was created in
spectral acceleration (Sa ) at the first period of the bridge in the skew OpenSees by Kaviani et al. [11] and modified by Wang et al. [57],
direction. Following an approach similar to that of ASCE 41-06 [55], which considered no RC deterioration. The columns are simulated with
the SRSS spectral acceleration (SaSRSS ) is first calculated from each pair force-based elements nonlinearBeamColumn in OpenSees with five in-
of rotated response spectra for the 21 records. Then the ground motion tegration points to control the numerical integration errors. Each in-
scale factor for each pair of horizontal records is calculated in such way tegration section includes 400 concrete fiber and 34 fibers for reinfor-
that the average SaSRSS in the interval 0.2T1T 0.1s to 1.5T1L 1.0s equals cing bars. Shear and torsional stiffness were also included using section
to 1.3 times the average spectral acceleration (Savg ) of the MCE spectra aggregator command in OpenSees. Linear-elastic beam-column ele-
for the bridge site in the same interval (in this case, ments were used to model the deck, with a mass and moment of inertia
SaSRSS = 1.3Savg = 0.933g ), where T1T and T1L denote the first period of based on the net area of the deck, in agreement with Caltrans [40]
vibration in the transverse and longitudinal direction of the bridge, requirements.
respectively. This scaling method captures the higher mode effects in In the longitudinal direction of the bridge model, five nonlinear
the longitudinal and transverse direction, as well as period elongation springs and the gap elements were used to simulate the passive backfill
effects due to inelastic behavior. response and expansion joints with the hyperbolic gap material in
As shown in Fig. 7, the method greatly reduces the dispersion of OpensSees. It considers the decrease of ultimate capacity of skew
response accelerations in the target period range of the far-field ground abutments, as the skew angle increases, in accordance with the ex-
motions. For example, the original dispersion at the transverse and perimental results from Rollins and Jessee [58] and Marsh [59]. Based
longitudinal bridge fundamental periods are 0.52 (T1T = 0.4 s) and 0.56 on these experiments, the spring stiffness variation was redefined for
(T1L = 0.66 s), and after scaling these dispersions are reduced to 0.38 each skewness. The largest stiffness variation along the skewed abut-
and 0.36, respectively. ment was 160% when the skew angle is 60°. Regarding the abutment
components in the transverse direction, the model includes piles, shear
2.6. Bridge model comparison keys and wingwalls by a zero-length element, with an overall stiffness
computed with Eq. (10) [1] and an elastic plastic backbone curve
To further compare the distributed plasticity model with the con- presenting the transverse system. A nonlinear spring with two stiffness
centrated plasticity model, THAs are performed for the straight bridge segments, acting only in compression, was implemented in the vertical
( = 0°) under different ground motions and intensity levels. Only one direction of the bridge abutment. The first slope represents the stiffness
horizontal ground motion is applied along the bridge transverse

Fig. 7. Response spectrum for the far-field ground motion set: (a) original; (b) scaled.

104
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Table 3
Periods of vibration of straight bridges.
Direction Concentrated plasticity (s) Distributed plasticity (s)

Longitudinal 0.64 0.66


Transverse 0.39 0.38
Vertical 0.29 0.29

attributed to a large decrease of the structural resistance of the col-


umns. The close seismic performance from the IDAs is an evidence of
the well-defined material degradation from concrete crushing, buckling
and tensile fracture of reinforcement in both the concentrated plasticity
model and the distributed plasticity model.

3. BRB retrofit and collapse capacity


Fig. 8. Vertical force-displacement backbone of the abutment.
The evaluation of collapse capacity for retrofitted bridges highly
of the elastomeric bearings (KB ), while the second stiffness (KSW ) has a depends on the BRB characteristics. Fig. 11 shows the geometric
larger value and represents the rigid behavior of the abutment stem parameters of a typical BRB configuration, including the connection
wall, as shown in Fig. 8. length (Lcon ), transition length (L tra ), core length (Lc ), and embedment
length (Lemd ). Design parameters values and failure modes from pre-
KSK Kpile vious experiments are summarized in Table 4.
K abut_T = + KWW = 2.71 × 105 kN/m
KSK + Kpile (10)
3.1. BRB failure mode
The retrofit scheme considers the addition of BRBs, as shown in
Fig. 2. This retrofit design between the steel gusset plates and the
To generate IDAs, the BRB failure drift was derived from two esti-
concrete components may require the implementation of steel rings
mators. First, the BRB failure drift was assumed to occur at the max-
around the columns, or alternatively, plates attached to the horizontal
imum axial strain that the BRB core plate is expected to reach. An in-
components at the face of the columns [18]. Fig. 9 presents the model
spection of previous experiments shows that the maximum BRB axial
configuration of the retrofitted skewed bridge, including BRBs at the
strain rarely exceeds the 3.5%, which generally equals to the air gap to
bents. The longitudinal periods of original bridges vary by only 3% due
BRB core yielding length ratio with a manufacture air gap limit of
to α variations, whereas the skew direction periods change as much as
75–100 mm (see Table 4). The air gap length is an important indication
15% as α varies from 0 to 54° [60]. Table 3 presents the fundamental
of BRB failure, since typical failures may occur after the BRB reaches to
periods of vibration of straight bridges from both numerical models.
this limit in past experiments, such as gusset plate buckling, steel case
The variation in corresponding periods is less than 5% for all cases.
bulging, and high-mode buckling of the core [62]. In this study, the BRB
axial strain limit is defined as 3.5%, which corresponds to a reasonable
2.7. Straight bridge IDA with concentrated and distributed plasticity upper bound approximation based on statistical results.
The second method correlates BRB axial strain to cumulative failure,
To extend the comparison to different seismic levels, IDAs were based on the recommendations from Andrews et al. [28]. An empirical
performed by applying ground motion records only in the transverse approach was developed to compute the BRB total CPD capacity, using
direction of straight bridges. As a common practice [4], SaSRSS is utilized a database of 76 BRB specimens, of which 34 failed during the test.
as intensity measure to capture the bidirectional seismic responses, and Andrews et al. [28] applied the maximum likelihood estimation
the lateral drift is selected as engineering demand. To obtain the IDA method, and considered a damage index based on deformation and
curves in Figs. 10, 21 far-field earthquake records were applied at dif- dissipated energy, to propose Eq. (11).
ferent seismic levels by scaling SaSRSS from 0 g to 16 g at an interval of
0.425 0.044
0.2 g. Individual IDA curves, as well as median, 16th and 84th per- TC = 2 21.20 AC LC 3.45 f u 1.46
( )
yc
centile curves, were generated for concentrated plasticity (Fig. 10a) and (A C) max (L C)max fy (11)
distributed plasticity (Fig. 10b) models, considering second-order P-Δ
where the BRB total CPD capacity (TC) is a function of the BRB core
effects. The statistical curves predict collapse capacities with a differ-
cross-section area, A C ; the BRB core length, LC ; as well as the statistical
ence less than 5% for both models, and the structural collapse can be
maximum BRB core area and core length from the 76 cyclic specimens,
(A C)max = 184.6cm2 and (LC) max = 472.1cm, respectively. The CPD also
depends on the BRB core yield and ultimate strength, fy and fu , re-
spectively; and the BRB core yield strain; yc . For the BRBs in this study,
the previous designed parameters [2] A C = 78.1cm2 , LC = 650cm are
used, as well as the typical BRB steel property, which are fy = 290MPa ,
fy
fu = 420MPa , and yc = E = 2 × 105MPa = 0.00144 .
290MPa

The BRB axial strain at failure was then derived for the evaluated
straight bridge, using this CPD criterion in THAs. For this purpose, IDAs
were carried out on the evaluated bridge model using the 21 far-field
records from FEMA P695. Fig. 12a shows the axial strains reached by
the BRB when the total CPD capacity is reached, according to Eq. (11).
Note that the axial strain of individual BRBs from Andrews et al. [28]’s
method may reach values in excess of 5%, but the median axial strain
from the 21 records is 3.1%, slightly smaller than the proposed air gap
Fig. 9. Retrofitted bridge model in OpenSees [2,8,25]. ratio upper bound of 3.5%. Fig. 12a also shows that spectral

105
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 10. Sa/g − θmax relationships for the straight bridge models with P-Δ: (a) concentrated plasticity; (b) distributed plasticity.

Fig. 12b).
Then, this study assumes a maximum BRB strain of 3.5% axial ratio,
which is a reasonable upper bound approximation based on statistical
results from the above two estimators. Because the BRB core length is
6500 mm, the estimated maximum core strain of 3.5% is 114 mm,
which is close to the maximum manufactured air gap limit. The BRB
brace system was modeled using a two-node link element. The BRB
components were modeled utilizing the Menegotto–Pinto (Steel02)
material in OpenSees [10]. This model includes isotropic and kinematic
hardening when using Pinching4 material in OpenSees software [69].
The BRB failure is implemented with MinMax material in OpenSees by
tracking the BRB axial core strain. In this study, the BRB system with-
stands 50% of the retrofitted bent’s shear force (DS-50BRB) [2], and
Fig. 11. Typical industrial BRB configuration (adapted from Xu and Pantelides after the BRB fails the lateral resistance is provided only by the concrete
[61]).
bent.

accelerations when TC is reached are equal or larger than 3 g.


For BRB qualification, AISC 341-10 [62] requires a minimum ex- 3.2. Effect of BRB failure on pushover and IDA curves
perimental CPD of 200 y at the end of the test, where y is the yield
length of the BRB core, which is equivalent to the LC × yc in Andrews Fig. 13a shows the pushover curves for the original three-column RC
et al. [28]’s method. Most BRB qualification tests, however, exhibit a bent in the straight bridge model, for two diagonal BRB components,
CPD capacity significantly larger than 200 y . Thus, IDAs were per- and for the retrofitted RC bent with the two BRBs. As observed, BRBs
formed in the evaluated straight bridge model to obtain the BRB axial fail once the 3.5% axial strain limit is reached, which is equivalent to a
strain at 200 y , 400 y , and 600 y , obtaining mean axial strains of lateral drift of 3.8%. At this point, it is assumed that the original RC
2.4%, 2.6%, and 2.95%, respectively. For the bridge model in this bent has to withstand all the seismic loads. To act as structural fuses,
study, a reasonable correlation between CPD and mean axial strain is BRBs should have a smaller yield displacement than that of the bare
reached when 600 y is used as the ductility capacity criterion (see bridge bent [17,18]. As a result, the BRB stiffness is chosen to have a
stiffness larger than that of the bare RC bent, concentrating seismic

Table 4
BRB design parameters and failure modes.
Test A c (mm2) L con (mm) L c (mm) Gap (mm) Gap to length ratio Max. axial strain Failure mode

Chou et al. [63] 1568 275 3836 160 4.4% 1% Gusset out of plane buckling
Chou and Chen [64] 3300 400 2650 140 5.2% 2.1% Tensile fracture
Xu and Pantelides [61] 11,936 3124 150 4.8% 4.1% Core end fracture, casing bulged
Xu and Pantelides [61] 11,936 3124 150 4.8% 3.5% Core mid-high tensile fracture
Xu and Pantelides [61] 6451 3886 150 3.8% 3.2% Weak axis buckling
Iwata and Murai [65] 2816 250 1251 46 3.6% 2.5% Core plate deforms and casing bulged
Iwata and Murai [65] 2288 250 1251 46 3.6% 3.0% stop Core plate strong axis deformation
Iwata and Murai [65] 1664 250 1251 46 3.6% 3.1% Tensile rupture of core
Matsui et al. [66] 2080 250 940 60 6.4% 3% Tensile fracture of core
Matsui et al. [66] 2080 250 940 60 6.4% 3% Core plate local buckling, casing deformation
Newell et al. [67] 7742 657 3391 150 4.4% 3.5% Bolts slip
Newell et al. [67] 17,419 657 3669 150 4.1% 3.2% Tensile fracture of core
Merritt et al. [68] 2451 577 4470 205 4.5% 2.4% Tensile fracture of core
Merritt et al. [68] 5380 552 4655 205 4.3% 1.8% Weak axis buckling of core plate
Merritt et al. [68] 11,529 622 4556 205 4.4% 2.5% Yielding of connection part
Zsarnóczay et al. [69] 2000 90 4.5% 3.5% Tensile fracture of core
Benzoni and Innamorato [70] 15,199 4721 200 4.21% 1.8% Tensile fracture of core

106
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 12. BRB failure under cumulative plasticity ductility prediction from: (a) Andrews et al. [28]; and (b) 600 y.

energy dissipation in the BRB yielding. Consequently, the BRB retro- frequency of skew bridges, and has a negligible influence on the long-
fitted bent shows a tri-linear pushover curve, with the first break point itudinal frequency of straight bridges, due to the orthogonality of the
representing the BRB yielding, and the second break point representing longitudinal and transverse directions in straight bridges. The long-
the yielding of the bare bent, Fig. 13b. itudinal periods of the original bridges vary by only 3% due to α var-
The IDAs of the distributed and concentrated plasticity models in iations, whereas the skew periods change as much as 15%, as α varies
the straight bridge realization with consideration of BRB retrofit are from 0 to 54°. This variation increases to 25% for retrofitted bridges
presented in Figs. 14a and 14b, respectively. The two models predict a because of the larger bent stiffness.
similar median collapse capacity of SaSRSS = 8g for the original and BRB
retrofitted straight bridges. Before reaching large lateral drifts, BRBs 4.2. Three-dimensional IDAs
significantly improve the bridge seismic performance, but BRBs do not
contribute to the lateral resistance for lateral drifts larger than 3.8%, Two horizontal time history accelerations from the 21 ground mo-
which corresponds to the 3.5% axial strain limit of assumed BRB tions in the FEMA P695 [54] far-field set, are applied to the bridges
failure. The IDA curves also show that the BRB retrofitting effect is along the transverse and longitudinal directions, with an increasing
significant until these large seismic drift demands are reached. There- intensity measure, SaSRSS , as well as the original accelerations applied in
after, the bare bent controls the collapse capacity of the retrofitted the vertical direction. The major principal ground motion direction,
bridges. As shown in Fig. 14, the BRB retrofitting effect decreases as the with higher Sa , is applied in the transverse direction [60]. Then, the
engineering demands increase, because of the earlier failure of BRB seismic accelerations are monotonically increased until complete col-
components. The DS-50BRB retrofit approach slightly increases the lapse is detected or the acceleration limit (SaSRSS = 16 g) is reached.
median collapse capacity of straight bridges with the concentrated
plasticity model, while the BRB has a smaller influence on the bridges
4.3. Collapse capacity of original skewed bridges
with distributed plasticity model.
The IDA curves in Fig. 15 represent the variation of the maximum
4. Incremental dynamic analyses for 3D skewed bridge models drift ratio demand as a function of the seismic level, SaSRSS (i.e., Sa/
with BRBs g − θmax relationships). The drift ratio is defined as the ratio of lateral
drift (displacement) to the total height of the bridge. The figure presents
4.1. Prototype skewed bridge models individual, median, 16th and 84th percentile IDA curves until complete
collapse is reached, as evidenced by the plateau in the curves. Com-
Different skew abutment angles (α) are utilized in this section, with pared to straight bridges (Fig. 15a), collapse occurs at smaller drift
respect to the longitudinal axis of α = 0° (straight), 18°, 36°, and 54°. ratios for bridges with large skew angles (Fig. 15b). For example, the
Table 5 presents the bridges’ fundamental periods of vibration for the median IDA curve becomes flat that indicates complete structural
original and DS-50BRB retrofitted bridge models under different skew failure or collapse, around 10% drift for the 54° skewed bridge, and at
angles. The BRB retrofit has a larger influence on the longitudinal approximately 14% for the straight bridge. The higher skew angle

Fig. 13. Pushover curves for DS-50BRB system, considering BRB failure limit: (a) curves up to 15% bent drift; (b) curves up to 2.5% bent drift.

107
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 14. Sa/g − θmax relationships for the straight bridge models with/without BRB: (a) concentrated plasticity; (b) distributed plasticity.

Table 5 computed for original and retrofitted bridges. The MAF of collapse is
Period of vibration of prototype skewed bridge models. obtained from numerical integration methods [34,71,72] that combine
Bridge Model 1st (Long. 2nd (Skew 3rd (Vertical Dir.)
fragility curves (FCs) obtained from IDA, and hazard curves (HCs)
Dir.) Dir.) calculated for the bridge location (i.e., Ripon, CA). In this study, FCs
were calculated in two ways with drift fragility distributed along the
= 0° (Straight), original 0.66 0.38 0.29 horizontal direction of IDA curves, and Sa fragility distributed along the
= 18°, original 0.66 0.39 0.29
= 36°, original 0.65 0.40 0.30
vertical direction of IDA curves (Fig. 18).
= 54°, original 0.68 0.44 0.33 First, the MAF of collapse was obtained based on a drift limit
= 0°, DS-50BRB 0.66 0.28 0.29 commonly associated to imminent collapse (i.e., 2.5% in this study
= 18°, DS-50BRB 0.64 0.31 0.29 [73]) to encompass the bridge performance at different seismic levels.
= 36°, DS-50BRB 0.58 0.32 0.29
Then, horizontal statistics of the maximum drift ratio were computed
= 54°, DS-50BRB 0.59 0.35 0.33
from the IDA curves. Thereafter, the MAF of complete collapse was also
Note: BRB lateral resistance contribution is 50% (DS-50BRB). computed using vertical statistics of individual collapse capacity results
(i.e., Sa ) under a drift limit associated to complete structural collapse
reduces RTR variability since the dispersion of individual responses at (i.e., approximately 13% in this study). In general, the latter approach
54° is reduced by about 10% (Fig. 7). However, the bridge skewness has is preferred, but the BRB retrofit’s influence on the seismic performance
a negligible effect on the median collapse capacity (Fig. 15c), since the may not be captured, if the BRB is assumed to fail at drifts significantly
bridge models with different skew angles have the same structural lower than the drifts leading to complete strength deterioration. For
properties, such as cross-section area and reinforcement, etc. comparison purposes, the imminent collapse limit of 2.5% drift is also
utilized in the vertical statistics, to assess the influence of BRB retrofit.
4.4. BRB retrofit effects in IDA
4.5.1. Mean annual frequency of imminent collapse – Horizontal statistics
This section evaluates the influence of BRB retrofit on the seismic In this section, the fragility function is derived based on “horizontal
performance of skew bridges. The BRB is assumed to reach complete statistics” of the maximum bent drifts at different Sa (Fig. 18), where the
failure when the BRB’s core axial strain reach 3.5%. Fig. 16a and b horizontal records are scaled based on the SRSS spectra from both di-
show the individual and statistical IDAs for the original and retrofitted rections, as in Section 2.5. The collapse FC (FC,ac ) is obtained from the
36° skewed bridges, respectively. As observed, the variability is smaller cumulative probability that the drift is larger than or equal to a pre-
for the bridges retrofitted with BRBs. For example, at a 12% drift, the defined imminent collapse drift (in this case, 2.5%) at monotonically
BRB components reduce the dispersion due to RTR variability by 23%. increasing seismic levels (i.e., Sa ), as shown in Eq. (12).
The influence of BRB retrofit on bridges with different skewness is
shown in Fig. 17, with the median, 16th percentile, and 84th percentile FC,ac (x) = P[D d c |Sa /g = x] (12)
curves. The benefits of BRB are evident before BRBs reach to the pre-
Thereafter, the distribution percentiles that the bent drift (D) ex-
defined failure lateral drift of 3.8%, but this performance improvement
ceeds the imminent collapse drift limit state (d c ) of 2.5% [73] are
is not uniform. BRB retrofit increases SaSRSS by 15%–20% at the 3.8%
summarized to calculate the fragility function with respect to increasing
lateral drift for low skew angles (i.e., 0° and 18° in Fig. 17a and 17b,
seismic levels.
respectively), while the retrofit can increase SaSRSS by 30%–40% for the
As the Sa increases, some of the drifts may be undefined because the
higher skew angles (i.e., 36° and 54° in Fig. 17c and 17d, respectively).
system starts to collapse under certain seismic records, as shown in
After BRBs fail, the retrofit benefits are greatly reduced, and the col-
Fig. 16. The median values for the fragility curve can be obtained using
lapse capacity is close for original and retrofitted bridges, with the
‘counted statistics’ [34], as long as the number of computed drifts is
difference of median values less than 5% for the bridges with low skew
equal or larger than the number of collapsed realizations. The approach
angles (i.e., 0° and 18°).
is then used to obtain the median and standard deviation of bridge
drifts at each seismic level. The median is a common central measure of
4.5. Fragility curves and mean annual frequency of collapse with distributed
dispersion for lognormal distributions, whereas the mean can be com-
plasticity
puted as follows:
To estimate the influence of BRB retrofit on the seismic performance 2
lnx
2
lnx
of skewed bridges, the mean annual frequency (MAF) of collapse is µ x = e µlnx e 2 = x50e 2 (13)

108
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 15. Sa/g − θmax relationships for bridges with different skew: (a) 0°; (b) 54°; (c) median comparison for all evaluated skew bridges.

where µlnx is the mean of the natural logarithm of drift values, x50 is the on the 16th and 50th [34,72] that is indicated in Eq. (15), as long as the
median of the data, and lnx is the standard deviation of the natural non-collapsed frames are sufficient to obtain the median value.
logarithm of drift values, which is calculated as [36]:
x50
x 84 lnx = ln
lnx = ln x16 (15)
x16 (14)
where x16 and x 84 refer to the percentile of the drift values under each Sa . Since the IDAs are computed for 21 ground motions in this study,
Because 21 ground motions are utilized in the current study, the 84th the median can be defined, as long as more than 11 realizations do not
percentile can be computed with the lowest 18 drifts, and the calcula- collapse at a given Sa .
tions are not affected by one or two collapse bridges. If more than three To obtain the MAF at imminent collapse, the seismic hazard, a (x)
records collapse, the standard deviation, lnx could be estimated based at the site (Latitude = 37.752852, Longitude = −121.142278, near

Fig. 16. Sa/g − θmax relationships for 36° skewed bridges: (a) Original; (b) Retrofit.

109
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 17. BRB retrofit effects of Sa/g − θmax relationships for bridges with different skew angle: (a) 0°; (b) 18°; (c) 36°; (d) 54°.

FCs, the bridges were subjected to 3D ground motion records, as pre-


sented in Section 4.2. Fig. 19 also shows the HC used in the calculations
and the probability distribution functions (PDFs) obtained from Eq.
(16). The MAF of imminent collapse for original and retrofitted bridges
are denoted as C_Orig and C_Retrof , respectively. In the straight bridge
model, the BRB retrofit reduces the MAF of the imminent collapse from
8.9 × 10-6 to 1.3 × 10-6, a reduction of approximately 85%. The PDF
curves also show that the Sa threshold for drift at imminent collapse
increases from 0.46 g in the original bridge to approximately 1.0 g in
the retrofitted bridge.
As the bridge skew angle increases, the MAF of imminent collapse
for the original and retrofitted bridges remains similar to that of
straight bridges, as shown in Table 6. As observed, the ratio of C_Retrof
to C_Orig decreases with the increase of the skew angle, which indicates
Fig. 18. Drift fragility and spectral acceleration fragility. a reduction of BRB retrofit effects for skewed bridges. Also, the Sa
threshold for potential imminent collapse decreases, as the skew angle
increase. For instance, the Sa thresholds are 0.46 g and 1.0 g for the
Ripon, California, USA) is reported in terms of Sa at the fundamental
original and retrofitted straight bridges, which are reduced to 0.38 g
period of the system in the skew direction, and Sa becomes the ground
and 0.6 g for the original and retrofitted 54° skewed bridges (Fig. 19).
motion intensity measure. The hazard curve (HC) required to compute
the MAF was obtained from the USGS [74] database for the funda-
mental period of the original system before retrofitted with BRBs in the 4.5.2. Mean annual frequency of collapse – Vertical statistics
skew direction, TSK 0.4s and 5% damping ratio. The MAF ( c ) of ex- In this section, the fragility function is the conditional probability of
ceeding the drift threshold associated to imminent collapse can be exceeding a certain complete collapse limit state for a given Sa . The
calculated by integrating the FCs and HCs at the site, as follows [36]. fragility function for collapse capacity is directly computed using ver-
tical statistics (see Fig. 18, and Ibarra and Krawinkler [34]), and com-
plete collapse can be approximated by computing the Sa associated with
c = FC(FC,a c )d | a (x)|
(16) this limit state, once all the IDA curves have reached the plateau (Sa,c ).
0
The bridge collapse capacity is directly obtained from numerical
The computed FCs for the bridge models with different skew angles models that account for strength and stiffness deterioration, see Section
before and after being retrofitted with BRBs are shown in Fig. 19, as- 4.2. The collapse capacity is computed at a drift of 13%, a drift in which
suming imminent collapse for a 2.5% drift ratio [73]. To generate these all realizations have collapsed and show straight plateau IDA curves.

110
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 19. HC, FC, and MAF of imminent collapse due to Sa = x at T = 0.5 s on bridges with different skew angles: (a) 0°, straight; (b) 54°.

Table 6 BRBs are assumed to fail at a lateral drift of 3.8%, and thereafter the
Summary of MAF of collapse with horizontal statistics. lateral seismic resistance is purely provided by the bridge bent, as in the
Skew angle Original Retrofit Ratio of retrofit to original
original configuration. Also, the vertical statistics at a drift of 2.5%
show that BRBs increase the Sa required to exceed a 2.5% drift limit by
0° 8.89 × 10 6 1.34 × 10 6 0.15 75%. This ratio is similar to that obtained from horizontal statistics
18° 8.76 × 10 6 1.44 × 10 6 0.16 (85%), although it is based on MAFs of exceedance one order of mag-
36° 6.89 × 10 6 1.21 × 10 6 0.18 nitude smaller, given that vertical statistics only capture the perfor-
54° 7.82 × 10 6 1.62 × 10 6 0.20 mance at a specific drift. A summary of MAF of collapse with ‘vertical
statistics’ is presented in Table 7.

Note that the BRB retrofit influence, particularly at early stages of


nonlinear behavior, may not be captured from vertical statistics since 5. Definitions, assumptions, and limitations
the BRB is assumed to fail at a drift of 3.8%. To better investigate the
BRB retrofit effect under ‘vertical statistics’, the Sa required to exceed 5.1. Discussion on column boundary definitions
the predefined 2.5% imminent collapse limit is also computed for ori-
ginal and retrofitted bridges. The collapse FC (FC,Sa ,c ) is the cumulative This section evaluates the effect of column bent boundary definition
distribution function (CDF) formed from individual seismic intensity at on the bridge collapse capacity. Previous studies [38,75] indicate that a
the collapse drift limit state (Sa,c ), as in Eq. (17). rebar pin moment-resistant model should be utilized as column to
footing boundary in lieu of the code specified “pure” pin connection
FC,Sa ,c (x) = P[Sa,c /g x] (17)
[38], see Fig. 2. In this section, the numerical model is modified to
The collapse FCs for straight bridges before and after BRB retro- investigate the influence of the rebar pin assumption on the bridge
fitting are shown in Fig. 20, using the same IDA curves presented in collapse capacity. Zero-length elements are assigned at the column base
Section 4.2. As observed, the MAF of complete collapse using ‘vertical to replace pin connections. The yield and ultimate moment capacity of
statistics’ at 13% drift is two orders of magnitude smaller than that rebar pin in the bridge case is 1933 kN-m, and the ultimate rotation is
computed from ‘horizontal statistics’. The main reason is that the ar- 0.127 rad, and are calculated based on the formulation presented in
bitrary drift limit of 2.5% imposed for the imminent collapse in the Caltrans [37] and Cheng et al. [75]. The parameters for the zero-length
‘horizontal statistics’ is smaller than the complete collapse drifts of the element are adopted from Mehraein and Saiidi [38].
individual IDA curve. Note that the BRB retrofit still reduces the MAF of The loading protocol used in Section 4.2 [53] is reapplied to the
complete collapse by 50%. This discrepancy reflects the fact that the bridge model that were discussed in Sections 2.3 and 2.6, but this time a

Fig. 20. HC, FC, and MAF of exceeding Sa on straight bridges under drift limit states of: (a) 13%; (b) 2.5%.

111
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Table 7 varies by only 5%. The results justify the assumption of a model with
Summary of MAF of collapse with vertical statistics. pure pin column connections, which is computationally stable for IDAs,
Drift Original Retrofit Ratio of retrofit to original and has been validated with experimental specimens under quasi-static
loading protocols. Moreover, the columns of the baseline bridge model
2.5% 2.85 × 10 7 6.73 × 10 8 0.24 are assumed to be pinned to the foundation, but the relative benefits of
13% 3.08 × 10 8 1.29 × 10 8 0.42 using BRBs to improve the bridge seismic performance are not expected
to be significantly different for other type of base column connections.

5.2. BRB connection detailing

The results presented in this study require adequate behavior of the


BRB-to-column connections. For RC bents, the BRB steel components
will be connected to existing RC columns that may exhibit different
seismic behavior, based on their design characteristics and aging de-
gradation, among other factors. For this reason, Bazaez and Dusicka
[18,22] and Al-Sadoon et al. [76] have proposed to attach the BRB to
connecting plates that transfer shear loads directly to the foundation.
These plates could use post-installed adhesive anchor rods to facilitate
the retrofitted method (Fig. 23a). Also, a steel plate jacket around
outside RC columns (Fig. 23b) could be used as a connection for BRB
retrofit [57]. Ideally, design information on the RC columns will be
available to verify that the composite action of the steel jacket and
concrete column is satisfactory. Alternatively, a parallel buckling re-
Fig. 21. Cyclic behavior of bridge bent model with different assumptions. strained braces frame (BRBF) could be constructed to resist lateral
seismic forces (Fig. 23c), as implemented in the retrofits of the Marriott
rebar pin model is considered. The cyclic behavior of the bridge with Library and the Wallace F. Bennett Federal Building in Salt Lake City,
the original pin column connection is very similar to that of the bridge UT.
with a rebar pin model, as observed in Fig. 21. The post-capping stiff-
ness is modified by less than 5%. For the complete bridge model, the 6. Conclusions
slight moment capacity reduction due to the pin assumption does not
lead to noticeable modifications on the expected structural damage. This study provides for first time a systematic approach to estimate
According to Fig. 2 dimensions, the pure bending capacity of the the deterioration and collapse capacity of skewed bridges retrofitted
footing is 10 time smaller than that of the column. As mentioned by with buckling-restrained braces (BRBs) by performing three-dimen-
Mehraein and Saiidi [38], concrete in compression at the column base sional incremental dynamic analyses (IDAs), based on a distributed
may develop a pair of forces with the steel in tension, but this effect will plasticity fiber model. This model was calibrated with the Ibarra-
decrease in the case study due to the presence of a thick filler material Krawinkler concentrated plasticity model to ensure that it can predict
that allows a certain rotation before the concrete is subjected to com- structural instability. Then, the effects of skew angle and BRB retrofit
pression. Also, under large seismic excitations the effect of the “pinned” scheme on the collapse capacity of reinforced concrete skewed bridges
connection moment capacity will decrease, given that the joint would were evaluated. The main findings are follows:
exhibit inelastic behavior earlier in the time history, further reducing
the potential moment that is transferred. (1) A larger bridge skew angle has a negligible effect on the median of
Furthermore, IDAs under 21 scaled GMs are generated for the 36° the maximum lateral drift ratio under IDAs. However, a larger skew
bridge model using the rebar pin model, see Fig. 22. A comparison of angle reduces record-to-record variability by about 10%. This re-
Figs. 22 and 16 shows that the adoption of the rebar pin model slightly duction is from the contribution of both horizontal ground motions,
reduces the collapse capacity for low drifts, likely associated to a and the lack of orthogonality between the longitudinal and the bent
change in the seismic demands. However, the median collapse capacity direction in skew bridges.

Fig. 22. Sa/g − θmax relationships for 36° skewed bridges: (a) with rebar pin model (b) results for model comparison.

112
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Fig. 23. BRB connections: (a) post-installed adhesive anchor rods; (b) steel plate jacket; and (c) outside steel frame.

(2) BRB failure limit was set at 3.5% axial strain, equivalent to a 3.8% scope of this study.
bent lateral drift ratio. This failure limit is based on typical BRB air
gap lengths between the concrete and steel core, and on a BRB Acknowledgements
cumulative plastic ductility failure method.
(3) Imminent collapse capacity was defined at a lateral bent drift ratio The authors are grateful to Mountain Plains Consortium (MPC) for
of 2.5%. Horizontal statistics showed that the bridge mean annual the financial support for this research under project MPC-421. The
frequency (MAF) at this limit state is reduced by 85% after the authors also thank Dr. Peyman Kaviani, Prof. Farzin Zareian and M. De
bridge is retrofitted. Bortoli for providing the original bridge model in OpenSees.
(4) Complete collapse was identified from flat IDA curves, which in-
dicate that the bent has lost its lateral strength resistance. Vertical References
statistics show that BRB retrofit increases the bridge’s collapse ca-
pacity at large drifts by only 10%–15%, given that BRBs are as- [1] Wang Y, Ibarra L. Pantelides C. Seismic assessment for retrofitted skewed reinforced
sumed to fail at a 3.8% lateral drift ratio. However, the BRB retrofit concrete bridges with buckling restrained braces. Proc 16th World Conf on
Earthquake Eng. Santiago: Chile; 2017.
reduces the bridge MAF of complete collapse by 60% because of the [2] Wang Y, Ibarra L, Pantelides C. Seismic retrofit of a three-span RC bridge with
contribution of moderate seismic events. buckling-restrained braces. J Bridge Eng 2016;21(11):04016073.
(5) The distributed plasticity model is an effective tool to perform [3] Ibarra L, Medina R, Krawinkler H. Hysteretic models that incorporate strength and
stiffness deterioration. Earthq Eng Struct Dyn 2005;34(12):1489–511.
three-dimensional IDAs up to the complete collapse limit state, as [4] Vamvatsikos D, Cornell CA. Incremental dynamic analysis. Earthq Eng Struct Dyn
long as adequate constitutive relationships are provided for the 2002;31(3):491–514.
concrete and steel materials. [5] Adam C, Ibarra L. Seismic collapse assessment. Encyclopedia Earthq Eng 2014:1–25.
[6] Altoontash A. Simulation and damage models for performance assessment of re-
(6) The MAF of imminent collapse is at the same order of magnitude for inforced concrete beam-column joints PhD. Dissertation Palo Alto, CA: Stanford
bridges with different skew angles. The influence of BRB retrofit on University; 2004.
the MAF of collapse slightly reduces with the increase of bridge [7] Haselton C. Assessing seismic collapse safety of modern reinforced concrete moment
frame buildings PhD. Dissertation Palo Alto, CA: Stanford University; 2006.
skew angle, while the seismic demand threshold (e.g., spectral ac-
[8] Lignos D, Krawinkler H, Whittaker A. Prediction and validation of sidesway collapse
celeration, Sa ) for potential imminent collapse increases as the skew of two scale models of a 4-story steel moment frame. Earthq Eng Struct Dyn
angle increases. 2011;40(7):807–25.
(7) The BRB retrofit influence can be captured by ‘horizontal statistics’, [9] Lignos D, Krawinkler H. Sidesway collapse of deteriorating structural systems under
seismic excitations. Rep. No. TB177. Palo Alto, CA: Stanford University; 2012.
commonly based on a drift limit commonly associated to imminent [10] McKenna F. Open system for earthquake engineering simulation. Berkeley, CA:
collapse (i.e., 2.5% in this study); and ‘vertical statistics’ of the MAF University of California; 2014.
of collapse. [11] Kaviani P, Zareian F, Taciroglu E. Seismic behavior of reinforced concrete bridges
with skew-angled seat-type abutments. Eng Struct 2012;45:137–50.
(8) Assumptions during modeling of the bridge bent may lead to dif- [12] Vamvatsikos D, Cornell CA. Incremental dynamic analysis with two components of
ferent conclusions. For instance, early structural collapse may occur motion for a 3D steel structure. In: Proc 8th National Conf on Earthquake Eng, San
with newly defined rebar pin model in lieu of code specified pin Francisco, CA; 2006.
[13] Fathieh A. Nonlinear dynamic analysis of modular steel buildings in two and three
connections. dimensions, Ph.D. Dissertation, University of Toronto, Toronto, Canada; 2013.
[14] Vamvatsikos D, Sigalas I. Seismic performance evaluation of a horizontally curved
The presented results are valid for BRBs implemented in bridges highway bridge using incremental dynamic analysis in 3D. Proc 4th European
workshop on Seismic Behavior of Irregular and Complex Struct. Greece:
designed according to recent codes, in which the operational limits
Thessaloniki; 2005.
need to be enhanced. The effect of BRBs on the seismic performance of [15] Billah A, Alam M. Seismic performance evaluation of multi-column bridge bents
older bridges is expected to be more significant, but it is not part of the retrofitted with different alternatives using incremental dynamic analysis. Eng

113
Y. Wang et al. Engineering Structures 184 (2019) 99–114

Struct 2014;62:105–17. [47] Lowes L, Moehle J. Seismic behavior and retrofit of older reinforced concrete bridge
[16] He R, Yang Y, Sneed L. Post-repair seismic assessment of RC bridges damaged with T-Joints. Rep No. UCB/EERC-95/09, Grand Forks, ND; 1995.
fractured column bars – a numerical approach. Eng Struct 2016;112:100–13. [48] Mazzoni S, Moehle J, Mahin S. Design and response of lower-level beam-column
[17] El-Bahey S, Bruneau M. Buckling restrained braces as structural fuses for the seismic joints in ductile reinforced concrete double-deck bridge structures. Berkeley, CA:
retrofit of reinforced concrete bridge bents. Eng Struct 2011;33(3):1052–61. Earthquake Eng Research Center; 1997.
[18] Bazaez R, Dusicka P. Cyclic behavior of reinforced concrete bridge bent retrofitted [49] Dhakal R, Maekawa K. Modeling for post-yield buckled of reinforcement. ASCE J
with buckling restrained braces. Eng Struct 2016;119:34–48. Struct Eng 2002;128(9):1139–47.
[19] Upadhyay A, Pantelides C. Comparison of the seismic retrofit of a three-column [50] Zhao J, Sritharan S. Modeling of strain penetration effects in fiber-based analysis of
bridge bent with buckling restrained braces and self centering braces. In: Proc reinforced concrete structures. ACI Struct J 2007;104(2):133.
Struct Congr 2017, Denver, CO; 2017. [51] Ameli MJ, Pantelides CP. Seismic analysis of precast concrete bridge columns
[20] Dong H, Du X, Han Q, Hao H, Bi K, Wang X. Performance of an innovative self- connected with grouted splice sleeve connectors. ASCE J Struct Eng
centering buckling restrained brace for mitigating seismic responses of bridge 2016;143(2):04016176.
structures with double-column piers. Eng Struct 2017;148:47–62. [52] Wu RY, Pantelides CP. Concentrated and Distributed Plasticity Models for Seismic
[21] Wei X, Bruneau M. Case study on applications of structural fuses in bridge bents. J Repair of Damaged RC Bridge Columns. ASCE J Compos Constr
Bridge Eng 2016;21(7):05016004. 2018;22(5):04018044.
[22] Bazaez R, Dusicka P. Performance assessment of multi-column RC bridge bents [53] Nguyen W, Trono W, Panagiotou M, Ostertag C. Seismic response of a hybrid fiber-
seismically retrofitted with buckling-restrained braces. Bull Earthq Eng reinforced concrete bridge column detailed for accelerated bridge construction.
2018;16(5):2135–60. Berkeley, CA: Earthquake Eng Research Center; 2014.
[23] Wei X, Bruneau M. Analytical investigation of bidirectional ductile diaphragms in [54] FEMA. Quantification of building seismic performance factors P695, Washington,
multi-span bridges. Earthq Eng Eng Vibr 2018;17(2):235–50. DC; 2009.
[24] Celik O, Bruneau M. Seismic behavior of bidirectional resistant ductile end dia- [55] ASCE/SEI. Seismic rehabilitation of existing buildings, Reston, VA; 2007.
phragms with buckling restrained braces in straight steel bridges. Eng Struct [56] Paulay T. A consideration of P-delta effects in ductile reinforced concrete frames.
2009;31(2):380–93. Bull N Z Ntl Soc Earthq Eng 1978;11(3):151–60.
[25] Lanning J, Benzoni G. Uang C. Using buckling-restrained braces on long-span [57] Wang Y. Seismic performance assessment of reinforced concrete skewed bridges
bridges. II: feasibility and development of a near-fault loading protocol. J Bridge retrofitted with buckling restrained braces. Ph.D. Dissertation, University of Utah,
Eng 2016;21(5):04016002. Salt Lake City, UT; 2017.
[26] Pantelides C, Ibarra L, Wang Y, Upadhyay A. Seismic rehabilitation of skewed and [58] Rollins K, Jessee S. Passive force-deflection curves for skewed abutments. J Bridge
curved bridges using a new generation of buckling restrained braces. Rep No. MPC Eng 2012;18(10):1086–94.
16-315. ND: Fargo; 2016. [59] Marsh A. Evaluation of passive force on skewed bridge abutments with large-scale
[27] Wei X, Bruneau M. Analytical investigation of buckling restrained braces’ applica- tests. Ph.D. Dissertation, Brigham Young University, Provo, UT; 2013.
tions in bidirectional ductile end diaphragms for seismic performance of slab-on- [60] Wang Y, Ibarra L, Pantelides C. Effects of ground motion incidence angle in re-
girder bridge. Eng Struct 2017;141:634–50. inforced concrete skewed bridge retrofitted with buckling restrained braces. In:
[28] Andrews B, Fahnestock L, Song J. Ductility capacity models for buckling-restrained Proc Struct Congr 2017, Denver, CO; 2017.
braces. J Constr Steel Res 2009;65(8):1712–20. [61] Xu W, Pantelides C. Strong-axis and weak-axis buckling and local bulging of
[29] Ibarra L, Bishaw B. High strength fiber reinforced concrete beam-columns with high buckling-restrained braces with prismatic core plates. Eng Struct 2017;153:279–89.
strength steel. ACI Struct J 2016;113(1):147–56. [62] AISC. Seismic provisions for structural steel buildings, Chicago, IL; 2010.
[30] Trejo D, Barbosa A, Link T. Seismic performance of circular reinforced concrete [63] Chou CC, Liu JH, Pham H. Steel buckling-restrained braced frames with single and
bridge columns constructed with grade 80 reinforcement. Rep No. Washington, DC: dual corner gusset connections: seismic tests and analyses. Earthq Eng Struct Dyn
FHWA-OR-RD-15-02; 2014. 2012;41(7):1137–56.
[31] Overby D, Kowalsky M, Seracino R. A706 Grade 80 reinforcement for seismic ap- [64] Chou CC, Chen SY. Subassemblage tests and finite element analyses of sandwiched
plications. Rep. No. CA16-2563. Sacramento, CA: California Department of buckling-restrained braces. Eng Struct 2010;32(8):2108–21.
Transportation; 2015. [65] Iwata M, Murai M. Buckling-restrained brace using steel mortar planks; perfor-
[32] Lowes L. Finite element modeling of reinforced concrete beam-column bridge mance evaluation as a hysteretic damper. Earthq Eng struct Dyn
connections. University of California, Berkeley. Berkeley, CA, vol. 1; 1999. 2006;35(14):1807–26.
[33] Naito CJ, Moehle JP, Mosalam KM. Experimental and computational evaluation of [66] Matsui R, Takeuchi T, Hajjar F, Nishimoto K, Aiken I. Local buckling restraint
reinforced concrete bridge beam-column connections for seismic performance. Rep. condition for core plates in buckling-restrained braces. In: Proc 16th World Conf on
No. PEER 2001/08, University of California, Berkeley, CA; 2001. Earthquake Eng, Santiago, Chile; 2017.
[34] Ibarra L, Krawinkler H. Global collapse of frame structures under seismic excita- [67] Newell J, Uang CM, Benzoni G. Subassemblage testing of corebrace buckling-re-
tions. Berkeley, CA: Pacific Earthquake Eng Research Center; 2005. strained braces (G series). Rep. No. TR-06/01, University of California, San Diego,
[35] Krawinkler H, Zareian F, Medina R, Ibarra L. Contrasting Performance-Based Design CA; 2006.
with Performance Assessment. Bled, Slovenia: International Workshop on [68] Merritt S, Uang CM, Benzoni G. Subassemblage testing of star seismic buckling-
Performance-Based Seismic Design; 2004. restrained braces. Rep. No. TR-2003/04, University of California, San Diego, CA;
[36] Deierlein G, Reinhorn A, Willford M. Nonlinear structural analysis for seismic de- 2003.
sign. Rep. NIST 10–917-5. Gaithersburg, MD: National Institute of Standards and [69] Zsarnóczay A, Dunai L, Kaltenbach L, Kallo M, Kachichian M, Halasz A. Type
Technology; 2010. Testing of Buckling Restrained Braces according to EN 15129–test reports.
[37] Caltrans. Bridge inspection records information system, Bridge No. 29 0318, Budapest, Hungary: Budapest University of Technology and Economics; 2011.
Division of Maintenance, Sacramento, CA; 2008. [70] Benzoni G, Innamorato D. Star seismic brace tests Mercy San Juan hospital project.
[38] Mehraein M, Saiidi M. Seismic Performance of Bridge Column-Pile-Shaft Pin Rep. No. SRMD-2007/05-rev2, University of California, San Diego, CA; 2007.
Connections for Application in Accelerated Bridge Construction, University of [71] Tsantaki S, Adam C, Ibarra L. Intensity measures that reduce collapse capacity
Nevada, Reno, Rep No. CCEER-16-01; Reno, NV, 2016. dispersion of P-delta vulnerable simple systems. Bull Earthq Eng
[39] Zhang J, Makris N. Kinematic response functions and dynamic stiffnesses of bridge 2017;15(3):1085–109.
embankments. Earthq Eng Struct Dyn 2002;31(11):1933–66. [72] Jalayer F, Cornell C. A technical framework for probability-based demand and
[40] Caltrans. Caltrans seismic design criteria version 1.1. Sacramento, CA; 1999. capacity factor (DCFD) seismic formats. Berkeley, CA: Earthquake Eng Research
[41] AASHTO. LRFD bridge design specifications. Washington, DC; 1998. Center; 2003.
[42] Haselton C, Deierlein G. Assessing seismic collapse safety of modern reinforced [73] OES (Office of Emergency Services). Vision 2000: Performance based seismic en-
concrete moment frame buildings. Berkeley, CA: Pacific Earthquake Eng Research gineering of buildings, Structural Engineers Association of California, Sacramento,
Center; 2007. CA; 1995.
[43] Fardis M, Biskinis D. Deformation capacity of RC members, as controlled by flexure [74] USGS. Unified hazard tool; 2017. See https://earthquake.usgs.gov/hazards/
or shear. Otani Symp 2003:511–30. interactive/.
[44] Mander J, Priestley M, Park R. Observed stress-strain behavior of confined concrete. [75] Cheng ZY, Saiidi M, Sanders S. A seismic design method for reinforced concrete
ASCE J Struct Eng 1998;114(8):1827–49. two-way column hinges. ACI Struct J 2010;107(5):572–9.
[45] Mander J, Priestley M, Park R. Theoretical stress-strain model of confined concrete. [76] Al-Sadoon Z, Palermo D, Saatcioglu M. Strengthening of deficient reinforced con-
ASCE J Struct Eng 1988;114(8):1804–26. crete buildings using a new type of buckling restrained brace. In: Proc 16th World
[46] Scott B, Park R, Priestley M. Stress-strain behavior of concrete confined by over- Conf on Earthquake Eng, Santiago, Chile; 2017.
lapping hoops at low and high strain rates. J ACI 1982;79:13–27.

114

You might also like