You are on page 1of 14

Engineering Structures 150 (2017) 176–189

Contents lists available at ScienceDirect

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

Comparison of curved prestressed concrete bridge population response


between area and spine modeling approaches toward efficient seismic
vulnerability analysis
Junwon Seo ⇑, Luke P. Rogers 1
Department of Civil and Environmental Engineering, South Dakota State University, United States

a r t i c l e i n f o a b s t r a c t

Article history: This paper presents two finite element modeling approaches for seismic evaluation of curved precast-
Received 23 September 2016 prestressed concrete (PSC) I-girder bridges and compares the results in a statistical and graphical manner.
Revised 13 June 2017 These approaches, including area and spine models, were applied to a simply-supported, curved PSC I-
Accepted 12 July 2017
girder bridge under an ensemble of 3D synthetic ground motions. Along with performing non-linear time
history (NLTH) analyses of the bridge to capture its separate seismic response, the efficiency of each
approach was evaluated with respect to execution time and each was compared. This comparison reveals
Keywords:
that the seismic responses that were computed at low computational cost from the spine approach are
Curved precast-prestressed concrete bridge
Spine model
reasonably analogous to those from the area model. Seismic fragility curves of a portfolio of curved
Area model PSC bridges using the spine approach are also created to assess their vulnerability and then compared
Seismic analysis to those gained from past studies. This comparison shows a reasonable agreement for the PSC portfolio.
Fragility Ó 2017 Elsevier Ltd. All rights reserved.
Efficiency
Ground motions

1. Introduction affect its seismic behavior and vulnerability [15], pertinent bridge
design codes, such as the American Association of State Highway
Bridges are considered the most vulnerable elements in the and Transportation Officials (AASHTO) Load and Resistance Factor
national transportation networks of the United States during an Design (LRFD) Seismic Bridge Design [1] have dealt with require-
earthquake [5,9,2]. In case of bridges designed without sufficient ments associated with the irregularities. Curved precast-
seismic detailing, it has been reported from past earthquakes prestressed concrete (PSC) I-girder bridges that have been
[7,19] that there has been significant damage on particular compo- regarded as one of the representative irregular bridge types have
nents of the bridges that were observed across regions in the Uni- frequently been constructed in both seismically active and moder-
ted States. Describing the probability of earthquake-induced ate seismic regions as the complexity of traffic flow transitions and
bridges experiencing different damage states has been vital to treat the efficiency of transportation network increases.
uncertainty related to their characteristics and incorporate ran- As stated above, several seismic vulnerability studies [7,18], for
domness in seismic excitations. A fragility curve has been com- straight PSC I-girder bridges and their retrofits [21,22] to seismic
monly used to quantify the probability of bridge damage or excitations have been performed. Specifically, Choi et al. [7] and
failure and vulnerability due to seismic loadings [33,20,23]. Nielson and DesRoches [18] examined seismic response of various
The majority of studies for seismic vulnerability assessments of straight PSC I-girder bridge configurations and types. It was found
bridges have focused on regular bridges in the form of fragility that simply supported straight PSC I-girder bridges exhibited seis-
curves [7,19,21,22,11,12,14]. However, there exist a significant mic fragility and potential damage under moderate ground
number of existing bridges with irregular configurations nation- motions. Padgett and DesRoches [23] proposed a methodology
wide. Since irregular configuration factors in a bridge unfavorably for the establishment of analytical fragility curves in part for
straight PSC I-girder bridges with different retrofit measures.
Though these studies have successfully assessed seismic vulnera-
⇑ Corresponding author. bility of straight PSC I-girder bridges, these findings are not appli-
E-mail addresses: junwon.seo@sdstate.edu (J. Seo), luke.rogers@hdrinc.com cable to curved PSC I-girder bridges due to their unique features
(L.P. Rogers).
1 and system mechanisms under seismic excitations.
Current address: Bridge E.I.T, HDR, 701 Xenia Ave. South, Suite 600, Minneapolis,
MN 55416, United States.

http://dx.doi.org/10.1016/j.engstruct.2017.07.033
0141-0296/Ó 2017 Elsevier Ltd. All rights reserved.
J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 177

Recent studies [29,28,30–34,3] demonstrated that steel bridges tensile strength of 1860 MPa and is 13 mm in diameter. Both
with irregular configuration factors such as radius of curvature straight and harped strands are employed. The concrete cast in
have more significant damage due to excessive torsional response place deck is 200 mm thick.
than regular bridges. For example, Seo and Linzell [29–31] have The bridge has the following attributes: the span was 23 m and
attempted to examine the seismic response and vulnerability of width was 9.8 m so that four girders were deemed appropriate; the
curved steel girder bridges using conventional three dimensional girders used 24 straight tendons and 8 harped tendons; the radius
(3D) finite element models. These models required more complex of curvature was 402 m; and an offset of 160 mm was developed at
computational solutions coupled with risk analysis for inherent the mid-span centerline as a result of the curvature and span
randomness in earthquakes in conjunction with uncertainty length. This with its design schematic can be illustrated in Fig. 1.
related to irregular characteristics than straight ones. It was The offset is the distance between the centerline and the arc chord
demonstrated that each curved steel bridge has distinctive charac- of curvature at mid-span. Eq. (1) will determine the offset as
teristics, leading to different seismic responses affecting its vulner- follows:
ability. The significant finding from the work [30,31] was that the
curved steel bridges produced fairly higher seismic vulnerabilities L2
O¼ ð1Þ
than those for straight ones due to their curvatures associated with 8R
the other parameters. Based upon the literature review, unfortu- where O, L, and R is the offset, span length, and radius of curvature,
nately significant studies related to seismic evaluation and vulner- respectively. Concrete cast-in-place diaphragms were 160 mm thick
ability of curved PSC I-girder bridges have not been found. To and placed at the midspan and endspan. The seat type abutment
better understand seismic behavior and vulnerability of curved was anchored by 8 cast-in-place concrete piles of 400 mm in diam-
PSC bridges at reasonably low computational cost using 3D finite eter. A 75 mm expansion gap was located at each abutment-deck
element models, the need of developing a new seismic modeling interface.
technique for such bridge populations that enables its efficient
seismic vulnerability assessment with variability in their bridge
characteristics is of significance. 3. Finite element modeling approach
The current study focuses on developing a finite element anal-
ysis modeling technique for seismic vulnerability assessment of Two modeling approaches, encompassing the area and spine
PSC I-girder bridges in a more efficient manner. To accomplish models, were presented on the bridge type using the program CSI-
the goal of this study, two 3D finite element modeling approaches, Bridge [8]. It is important to note that area and spine modeling is a
including a spine model and an area model, were developed and general term that can describe many different models. The models
applied to a curved two-lane, simply supported PSC I-girder bridge. described in this study were specific to the finite element modeling
The responses of both models of the bridge were compared statis- approaches used for the seismic response and fragility investiga-
tically and graphically. The bridge was designed using current tion of curved PSC bridges and were categorized into area and
AASHTO codes [1] and PCI (Precast/Prestressed Concrete Institute) spine as appropriate. The spine modeling approach is one that
practices [24]. Nominal materials and components for bridge con- models a superstructure with beam elements, while the area
struction were selected for use in the design. The 3D models were model approach use plate elements or shell elements to idealize
loaded with an ensemble of synthetic ground motions having two it. The following subsections describe each approach for the super-
horizontal and one vertical acceleration components. These ground structure and the modeling for the substructure with bearings for
motions varied in magnitude and intensity to make it possible to both approaches.
run a broad spectrum nonlinear time history analysis of seismic
response. The effect of model approach on the seismic response 3.1. Area model for superstructure
is also examined, along with comparing efficiency of the analysis.
To validate the possibility of creating seismic fragility curves, The area modeling approach is the more complex of the two
component- and system-level fragility curves for a suite of 15 approaches in the current study. The superstructure model is con-
curved simply supported PSC I-girder bridges that were also structed using shell and frame elements. Prestress tendons repre-
designed following the AASHTO and PCI specifications were devel- senting the Gr. 270, 7-wire strands were modeled using frame
oped using a spline model and compared to those developed in elements located within each girder. Tendon prestress force is
past studies of straight ones. 1100 MPa after all losses, including elastic shortening, steel relax-
ation, creep, and shrinkage. Both harped and straight tendons were
modeled and placed in the correct position relative to the girder
2. Structural design of curved PSC girder I-girder bridge according to the structural design.
Each girder was represented by frame elements with the cross
A curved PSC I-girder bridge was initially chosen and designed to section of a Type III girder that span between the supports. The
examine its seismic response and susceptibility and to explore the concrete compressive strength of the girder is 48 MPa. Shear rein-
differences between area and spine models. The bridge design com- forcement consists of #10 (#3 imperial) stirrups that were placed
plied with the [1] LRFD Bridge Design Specifications [1] following appropriately according to structural analysis. These frame ele-
the procedure outlined by the Precast/Prestressed Concrete Insti- ments were assigned the correct cross section and material proper-
tute Bridge Design Manual [24]. It was assumed that the bridge is ties of the girders specified during the designing process. The
located in the central United States that would be considered Seis- bridge deck was idealized using shell elements. Maximum length
mic Zone 1 where design considerations for seismic performance of a particular shell element is 3.05 m. Increasing this length
are minimal. These requirements are described in the AASHTO decreases the number of joints and vice versa. The deck section
Specifications. The abutment type for these bridges is the seat- shell elements represent a 200 mm slab thickness consisting of
type abutment. Integral abutments are generally not favorable for 27 MPa concrete. Note that a shell elements are three or four-
curved bridges [16]. Elastomeric pads each with two embedded node area objects used to model membrane and plate-bending
steel dowels were selected for use in both fixed and expansion bear- behavior and are useful for simulating bridge deck systems [8].
ings. The AASHTO I-girder type III was used based on established Diaphragms were also modeled using shell elements with cor-
design requirements. Each girder is prestressed with strands of ner nodes at adjacent girder frame elements; thus, with four gird-
178 J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189

(a) Top view

(b) Elevation

(c) Cross-section at mid-span

At Midspan At Endspan
(d) Girder Detail
Fig. 1. Bridge plan and details.
J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 179

ers for the bridge, there are three diaphragm shell elements, one study for the piles. Foundation springs under the abutment were
placed within each of the spaces between girders. The position of modeled accounting for both passive and active earth resistance
the diaphragms was determined according to the designed bridge via use of multi-linear springs. This was achieved by defining
and placed accordingly to connect adjacent girders. A single two springs set in parallel, one for the passive pressure and the
200 mm thick concrete diaphragm was used on this bridge and is other for the piles. Passive resistance is encountered when the
placed at the midspan. Fig. 2(a) shows a plan view of the area abutment pushes into the backfill. Passive resistance was calcu-
model of the superstructure, and Fig. 2(c) illustrates an elevation lated using techniques from previous research [6,17]. The initial
view of the girder with tendons in this model. stiffness of 23 kN/mm for passive resistance was obtained. Passive
resistance increases as deformation increases; thus, a multi-linear
3.2. Spine model for superstructure spring was used to model this behavior. Active resistance is pro-
duced by the foundation piles. In the transverse direction, the pile
The spine model is the most common approach used in past resistance is the only factor that should be accounted for. The effect
studies of seismic vulnerability of bridges [6,17,27]. This approach of wing walls was ignored as in past studies [7,18].
has less complexity compared to the area one but at the cost of The abutments were modeled using frame elements having
having fewer joints to measure responses. The spine model for appropriate dimensions that were determined from the structural
the superstructure is primarily made up of frame elements. The design. Nominal material properties of the concrete were assigned
entire bridge deck and all of the girders are modeled using a single to the abutment frame elements in the model. The compressive
frame element; this portion of the model is the most dissimilar strength of the concrete used in the abutment is 27 MPa. Main flex-
from the area model. The spline modeling approach computes ural reinforcement of the abutment consists of eight #22 bars
frame element properties such as stiffness by identifying the mate- (#7 bars imperial). Depth of the seat area of the abutment is
rial properties, dimensions, and locations of girders and deck sec- 910 mm. Height of the backwall is 1.37 m to accommodate the
tions. As a result of the superstructure modeling identification, height of the Type III girder and bearing.
the element has attributes such as axial stiffness, bending stiffness,
torsional stiffness, and mass similar to that of the area model. It
4. Seismic analysis procedure
should be noted that the tendons were modeled in the same way
of the area model. This modeling approach greatly reduces analysis
Seismic analysis procedure used for this study is detailed in the
time and computational expense due to the fewer number of
following subsections. This consists of the selection of synthetic
joints. A disadvantage of spine models is the limited analysis of
ground motions and non-linear time history analysis (NLTH) for
membrane effects as compared to area models. Fig. 2(b) displays
both modeling approaches.
the plan view of spine model superstructure.

3.3. Model for substructure with bearings 4.1. Synthetic ground motions

Substructure components for both approaches were modeled Since useful and widely varying real ground motion time histo-
using links representing springs and frames available in the CSI ries are not readily available, 300 synthetic ground motions across
program. A schematic of the bridge showing the bearing, founda- the regions of interest, including the Central United States such as
tion, and abutment can be seen in Fig. 2(c). The detailed substruc- Memphis, were obtained from the University of New York at Buf-
ture model can be seen in Fig. 2(d). All the substructure falo [10]. As shown in Fig. 3 for a Peak Ground Acceleration
components with the bearings are the only elements that were (PGA) histogram for the synthetic ground motion, the ground
modeled in the same way between both area and spine models. motion pool has a broad spectrum of PGA values, and each includes
The bearings were represented by two separate spring elements transverse, longitudinal, and vertical acceleration components.
located at the end of each girder. One spring was used for the elas- An approach of ground motion application used for this study
tomeric pad and one for the embedded steel dowels. Appropriate was decided based upon the recommendations in the past work
values for horizontal bearing stiffness were obtained from equa- [29–31]. It was recommended that the horizontal ground motions,
tions presented in past studies [6,7,17,18]. The 25 mm thick elas- including longitudinal and transverse components, be applied to
tomeric pad effective stiffness was determined to be 8.2 kN/mm. curved bridges. Note that past studies [7,35] used ground motions
An initial stiffness for the steel dowels within the bearing is containing only the horizontally longitudinal and transverse
92 kN/mm. Stiffness for the vertical axis on the bearing allowed ground motions. The seismic fragility data of the bridges used in
for free movement upward and restricted movement below the the past work [6] was directly compared to those from this study.
resting position. This was modeled using gap elements available Additionally, in the current study, it was preferred to incorpo-
in the program. This simulates no vertical restraint and allows rate the vertical accelerations during the nonlinear time history
the superstructure to lift-off from the bearing and abutment. In analysis since curved bridges have varying geometry that could
this way, actual conditions were replicated where the self-weight potentially exaggerate the effects of vertical ground motions. Nine
of the superstructure is the only significant resistance to vertical ground motions covering a wide array of PGAs were randomly
lift-off from the support. Fixed bearings have a gap of 5 mm before selected for use in this study. These ground motions had PGA
the dowel is activated while the expansion bearings have a gap of ranges of 0.1 g (minimum value), 0.5 g (median value), and 1.0 g
25 mm before the dowel is activated. Alongside the bearings, (maximum value). Fig. 4 provides the spectral acceleration graphs
impact elements were modeled using multi-linear springs to sim- representing each components of the ground motions. Fig. 4(a)
ulate the impact of the superstructure with the abutment backwall shows the spectral acceleration verse period for the longitudinal
during seismic events. A 75 mm expansion gap separates the abut- (x) acceleration where a maximum of 3.6 g, a minimum of 0.4 g,
ment from the superstructure. When this gap closes, impact occurs and a median of 1.75 g is shown. Fig. 4(b) displays the accelera-
and the spring is activated. Values for these springs were deter- tions for the transverse (y) direction. Here, the maximum spectral
mined using formulas from past research [17]. acceleration is 3.8 g, a minimum of 0.45 g, and the median of 1.7 g.
The foundation piles were modeled using spring elements. Stiff- Lastly, in Fig. 4(c), the vertical (z) spectral accelerations are shown
ness was obtained from past studies for cast-in-place concrete where the maximum is 1.95 g, a minimum of 0.5 g, and a median of
piles [7,18,4]. An effective stiffness of 7 kN/mm was adopted in this 1.6 g.
180 J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189

Shell Element

X (a)

Beam-Column Frame Element

Typ. Girder line

X
(b)

Detail (d) and (e).


Tendon Elements Rigid Link
Force (kN)

k = 5.4 kN/mm

Z Displacement (mm)
Elastomeric Pad
X

k2 = 2.1 kN/mm
Force (kN)

Force (kN)

Force (kN)
Force(kN)

k1 = 10.9 kN/mm
k1 = 23 kN/mm

k = 93 kN/mm k2 = 3.9 kN/mm


k = 93 kN/mm
k3 = 1.6 kN/mm
Displacement (mm) Displacement (mm) Displacement (mm) Displacement (mm)

Fixed Steel Dowel Expansion Steel Dowel Pile Passive Soil Resistance

(c)

(d)
Fig. 2. 3D finite element model details: (a) area model superstructure plan view; (b) spline model superstructure plan view; (c) elevation view with spring properties; and (d)
model substructure detail.
J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 181

4
3.5

Spectral Acceleraon (g)


3
2.5
2
1.5
1
0.5
0
0 1 2 3 4 5
Fig. 3. PGA histogram for the ground motions. Period (s)
(a)
4.2. Non-linear time history analysis
4
Non-linear time history analysis was used to capture seismic
3.5
responses of the two modeling approaches applied to the bridge.

Spectral Acceleraon (g)


Although this seismic analysis technique tends to be the most 3
computationally expensive, this analysis is considered one of the
most reliable methodologies available in earthquake engineering 2.5
[33]. As such, there have been many researchers that have used a 2
methodology based on NLTH analysis to determine nonlinear
bridge responses resulting from seismic excitations. The NLTH 1.5
analysis procedure is outlined in Fig. 5. This procedure is done a
1
number of times for each bridge, once for each ground motion time
history. The two modeling approaches were considered and the 0.5
ground motions are paired with the bridge models and a non-
0
linear time history analysis was performed. For each simulation, 0 1 2 3 4 5
the peak seismic responses for key bridge components such as
Period (s)
bearing deformation were collected. In addition to peak responses,
the complete time history of loading and deformation of certain (b)
components was extracted from the results to create hysteresis.
4
5. Seismic analysis results 3.5
Spectral Acceleraon (g)

The efficiency of each approach that was evaluated with respect 3


to execution time is detailed herein. Seismic responses of critical 2.5
components were captured NLTH analyses of the bridge between
the two approaches are also summarized in the following subsec- 2
tions. Included in the critical components were the bearings, gird-
1.5
ers, and abutments.
1
5.1. Analysis efficiency
0.5
The main advantage of the spine model over the area model is 0
the reduction in complexity. This analysis was performed with 0 1 2 3 4 5
all six degrees of freedom (DOFs) enabled, although there is the Period (s)
option to disable any number of them. As can be seen in Table 1, (c)
the number of joints in the spine model is 33% lower than in the
area model. This is the core factor that affects other factors and Fig. 4. Spectral acceleration of selected ground motions in different orthogonal
directions: (a) longitudinal direction; (b) transverse direction; and (c) vertical
ultimately determines the number of DOFs and equations of equi-
direction.
librium. As noted, the reduction in frame elements by 46% from the
area to spine model is a direct result of the lack of shell elements
within the spine model. Computational time is the result of the using a spine model is the lower number of equilibrium equations.
program creating and solving matrices based on the constraints From the area model to the spine model, the reduction is on the
and DOFs. Due to the lower element and joint count, the number order of 48%. Similarly, the reduction in nonzero stiffness terms
of constraints is lower in the spine model. This also lowers the is around 46%.
number of constraint equations by 24% and lowers the Master The aforementioned reduction in number of joints, elements,
DOF significantly by 76% from the area model to the spine model. constraints, and DOFs ultimately reduces computational time by
Even after DOF reductions due to coupled constraints/restraints, reducing the number of equilibrium equations. This reduction is
the final Constraint Master DOF is 68% lower in the spine model time on the order of 7% for this size of seismic analysis. The
than in the area model. The greatest benefit that can be seen when computer used for this analysis utilized a 32-bit operating system,
182 J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189

Abutment
Acceleration (g)

Bearing

Girder

Time (sec)
Time (sec)
Fig. 5. Non-linear time history analysis procedure.

Table 1 4 GB of ram, and a 3.6 GHz dual-core processor. These specifications


Number of joints, elements, and DOFs for area versus spine models. directly impact analysis speed which is timed by the program CSi-
Area Spine Percent Bridge during an analysis run. The size of the analysis performed for
Difference (%) the designed bridge would be considered small in comparison to
Number of Joints 637 424 40 typical analysis with larger bridge models and more ground motion
Number of Joints with Restraints 48 48 0 time histories as in past studies of bridge fragility [6,17,27]. Due to
Number of Frame/Cable/Tendons 606 330 59 increased complexity of larger models, it can be expected that the
Number of Shells 69 0 200
time savings when using the spine model over the area model
Number of Link/Supports 216 219 1.3
Number of Constraints 876 850 3.0 would be significant for seismic vulnerability assessment of a num-
Number of Coupled Constraint Equations 1096 836 27 ber of bridges with different structural characteristics.
Constraint Master DOF before Reduction 1264 306 120
Coupled Constraint/Restraint Master DOF 160 12 170
Constraint Master DOF after Reduction 1104 294 120 5.2. Seismic responses of critical components
Total Number of Equilibrium Equations 2598 1344 64
Number of Non-Zero Stiffness Terms 60,045 32,460 60 5.2.1. Bearings
Analysis Time (min:sec) 46:01 43:03 7
The elastomeric bearings, both fixed and expansion types,
undergo displacements up to 33 mm as a result of the ground

Fig. 6. Bearing displacements: (a) vertical; (b) transverse; (c) longitudinal; and (d) percent difference.
J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 183

Fig. 7. Displacements at midspan: (a) vertical; (b) transverse; (c) longitudinal; and (d) percent difference.

motions. These displacements are relative, meaning the difference 5.2.3. Abutments
between the bearing spring element ends that connect the girder The abutment displacements displayed are the average of both
to the abutment. In Fig. 6, it can be seen that the difference in abutments. The difference between these was minor, less than 5%,
results between the two model approaches varies depending on so the average is an accurate representation. Fig. 8 shows these
the direction of displacement. Fig. 6(a) displays the vertical results for the abutments. Fig. 8(a) shows the vertical displace-
deformation where spine model reaches 2 mm while the area ments where the spine model reaches 1.5 mm and the area model
model reaches 1 mm. Fig. 6(b) shows the transverse displacements. reaches 0.6 mm. Fig. 8(b) shows transverse displacement, this is
These are the most similar across model approaches where the the most similar between model approaches. A maximum of
maximum reached is around 11 mm for both model approaches. 35 mm is reached for both models. In Fig. 8(c) the longitudinal
Fig. 6(c) shows the longitudinal movements where the area response is shown, these values reach 3.5 mm for the area model
model reaches a greater displacement at 33 mm verses the spine and 2.75 mm for the spine model. These values are low mainly
model at 25 mm. In sum, the percent differences are displayed in due to passive resistance provided by the abutment backfill.
Fig. 6(d). It can be seen that the greatest difference lies in the Fig. 8(d) shows the percent differences. Again, the transverse direc-
vertical direction at an average of about 90%, seconded by the tion is most similar between models with an average percent dif-
longitudinal direction at 40%. The transverse direction is the most ference of less than 15%. Differences in the longitudinal direction
consistent between models with a percent difference less than 5%. are about 90%. In the vertical direction, it is about 35%. As stated
Discrepancies in bearing displacements between the approaches before, overlapped shell elements caused the discrepancy in abut-
are caused by overlapped shell elements in the deck section of ment displacements between models.
the area model.
5.2.4. Hysteric analysis of critical components
Fig. 9 shows hysteresis loops for the various components of the
5.2.2. Girder
bridge for both area and spine modeling approach. Included from
Global displacements were measured at the mid-span of
the both approaches are the elastomeric pad, steel dowels, and
girder 2 on the centerline. Fig. 7 displays the results for all
piles. In Fig. 9(a) the pad in the area model reaches a value of
component directions at the mid-span between the two model
25 mm at 75 kN where as in the spine model the pad reaches
approaches. Fig. 7(a) shows the vertical displacements where
15 mm and 75 kN. A difference of 50%. Fixed steel dowels shown
the area model reaches a maximum of 7 mm compared to the
in Fig. 9(b) are similar reaching the yielding force of 105 kN and
6.5 mm of the spine model. Fig. 7(b) shows movement in the
then fracturing, as indicated later in the hysteric loop by the exces-
transverse direction where the results are almost identical
sive displacement with zero force. Fig. 9(c) shows the cast-in-place
between model approaches, a maximum of about 48 mm is
pile hysteresis from the spine model and the area model. The max-
reached. In Fig. 7(c), the longitudinal displacements are shown.
imum force is 140 kN at about 23 mm for the area model and
The area model reaches 38 mm while the area model reaches
21 mm at 125 kN for the spine model. The percent difference in
30 mm. Finally, in Fig. 7(d) the percent differences are displayed
displacement is 9% and the difference in force is 11%.
showing that the vertical and transverse directions are very con-
sistent between modeling approaches with a percent difference
of less than 5%. The longitudinal average difference is around 5.2.5. Model comparison to past studies
40%. Again, the discrepancy between models is due to the over- Past bridge seismic analysis studies that incorporated the
lapped shell elements in the deck causing a higher mass than straight single span PSC girder bridge inventory include Choi [6]
the spine model. and Nielson [17]. The material properties are consistent across
184 J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189

Fig. 8. Abutment displacements: (a) vertical; (b) transverse; (c) longitudinal; and (d) percent difference.

the studies and procedure for determining bearing, pile, and pas- parameters, each with three possible values. This procedure is out-
sive action stiffnesses. The past studies used spine models, and lined by the past work [30]. The three parameters were bridge span
the results are compared to those resulting from the spine model length, bridge width, and offset. Three values each for the span
of the current study. The geometry and significant results are sum- length and bridge width are selected based on literature review
marized in Table 2a and b comparing bridges in past studies to the and engineering judgment. The offset parameter is taken as the
bridge used herein. Note that the past studies summarized the recommended value from the PCI [24]. This procedure was used
results of a typical bridge under a 0.6 g PGA ground motion for to attempt to include as many combinations as possible while
an example. The bridges used in the past work by Choi [6] and maintaining a feasible amount for study.
Nielson [17] are comparable to the bridge studied herein, although
the Choi bridge has the longer span length compared to the current 6.1. Limit states
study bridge. It can be seen that the mid-span displacement of the
current study bridge girder in the transverse direction is larger Limit states are needed to quantify damage during a NLTH anal-
than that of the girder resulting from the Nielson [17] study. ysis and are used to create seismic fragility curves. These are
There has been a discussion as to whether or not the transverse defined in terms of displacements for abutments and bearings,
direction is vulnerable to seismic loadings and results of studies while a curvature or drift is used for columns. These states in the
have been somewhat conflicting [6,17]. A contributing factor to current study follow damage state (i.e., slight, moderate, extensive
the high transverse abutment movement in the current study is and complete) definitions developed by FEMA for the HAZUS-MH
that there is no accounting for the wing wall resistance. Accounting loss assessment package [36]. As in recent studies [25,30], the limit
for this would lower these displacements. Also, longitudinal move- states developed by Nielson [17] are implemented. These are a
ment is slightly lower due to the dimensions of the elastomeric combination of physics based and survey based values. To blend
bearing used on the bridge. Since there are four girders for the cur- these values for each damage state, the Bayesian approach was
rent study, each bearing is required to be larger, and therefore stif- used which is outlined by Nielson [17]. The displacement values
fer than a bearing on a five girder bridge that was used for the for the limit states used in the current study are shown in Table 4.
Nielson study [18]. It should be noted that because the vertical displacement limit
states were included in the past work, fragility analyses with seis-
6. Seismic fragility analysis mic response quantities resulting from vertical ground motions
were not unfortunately made in this study.
A suite of 15 curved PSC bridges having various radii of curva-
ture is selected for evaluating their seismic vulnerability in the 6.2. Probabilistic seismic demand models and fragility curve
form of fragility curves. These bridges are modeled using the spine generation
model approach and the procedures described above. All bridges
mentioned herein are single span with varying span length, width, To develop fragility curves, probabilistic seismic demand models
and offset. A summary of the bridges included in this fragility study (PSDMs) must be generated using displacement data from the NLTH
is shown in Table 3. The various geometry and component charac- analysis. This assumes a lognormal distribution and helps correlate
teristics are given. The 15 bridge configurations were obtained demand to intensity measure [7]. The intensity measure is taken as
using the Central Composite Design (CCD) method based on three a PGA for this study. A more detailed description is given in past
J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 185

Fig. 9. Hysteresis loops under median PGA ground motions: (a) elastomeric pad; (b) steel fixed dowel; and (c) cast-in-place concrete piles.

work [6,17]. Once PSDMs have been created, the component fragi- first-order reliability theory having upper and lower bounds. This
lity curve is generated for each bridge component at each damage represents the system-level probabilities of exceedance and is cal-
state using Eq. (2), which have been used in various fragility studies culated by using Eq. (3) [7,13].
[23,29,26,15,34]. This fragility generation approach has been con- Y
m
m
sidered the most straightforward and common way to estimate max½PðF i Þ 6 PðF sys Þ 6 ½1  PðF i Þ ð3Þ
i¼1
seismic fragility behavior of structural components. i¼1
0   1
SD where PðF sys Þ is the cumulative probability, PðF i Þ is the component
B ln SC C
Pf ¼ U@qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiA ð2Þ probabilities. The fragility curve generation procedure is further
2
bd þ bc 2 detailed in past studies [7,23,13].

where Pf is the exceedance probability, Sc is the median value of the 6.3. Comparison of seismic fragility curves
structural capacity defined for the damage state, bc is the dispersion
or lognormal standard deviation of the structural capacity, Sd is the A set of component fragility curves was generated for the single
seismic demand in terms of a chosen ground motion intensity span curved PSC bridge class at different damage states following
parameter, bd is the logarithmic standard deviation for the demand Eq. (2). Component level fragility curves for the current study are
and UðÞ is the standard normal distribution function. Once compo- displayed in Fig. 10 for the fixed bearing, expansion bearing, abut-
nent level fragility analysis is complete, the component-level fragi- ment with fixed bearings, and abutment with expansion bearings.
lity data were used to create a system level fragility curve following These component fragility curves were then compared to those of
186 J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189

Table 2
Comparison between past studies and current study: (a) bridge characteristics and (b) seismic responses for critical components.

Bridge Characteristic Choi [6] Nielson [17] Current Study


(a)
Span (m) 34.4 18.3 21.3
Width (m) 8.68 9 9.1
No. Girders 5 5 4
No. Piles/Abutment 31–74 8 8
PGA (g) 0.6 0.65 0.5
Critical Component Seismic Response
Choi [6] Nielson [17] Current Study
(b)
Bearing (mm) Long. 100 28 18
Trans. 100 5 7
Vert. NA NA 1
Abutment (mm) Long. 3 <7 22
Trans. 3 <7 3
Vert. NA NA 0.6
Girder at mid-span (mm) Long. NA 30 22
Trans. NA 7 31
Vert. NA NA 7

Table 3
Single span PSC bridge inventory.

Bridge Geometry Beam Abutment


Span (m) Offset (m) Curve Radius (m) Width (m) Beam Type No. of Beams Total Tendons No. Piles Back wall Height (m)
1 21.3 15.2 373.6 9.1 I-45 4 32 8 1.4
2 21.3 15.2 373.6 18.3 I-45 7 32 10 1.4
3 21.3 30.5 186.9 12.2 I-45 5 32 8 1.4
4 21.3 45.7 124.7 9.1 I-45 4 32 8 1.4
5 21.3 45.7 124.7 18.3 I-45 7 32 10 1.4
6 33.5 15.2 922.3 12.2 I-63 5 50 10 1.8
7 33.5 30.5 461.3 9.1 I-63 4 50 10 1.8
8 33.5 30.5 461.3 12.2 I-63 5 50 12 1.8
9 33.5 30.5 461.3 18.3 I-63 7 50 16 1.8
10 33.5 45.7 307.6 12.2 I-63 5 50 10 1.8
11 48.8 15.2 1951.3 9.1 I-72 7 68 14 2.0
12 48.8 15.2 1951.3 18.3 I-72 13 68 24 2.0
13 48.8 30.5 975.8 12.2 I-72 10 68 18 2.0
14 48.8 45.7 650.6 9.1 I-72 7 68 18 2.0
15 48.8 45.7 650.6 18.3 I-72 13 68 24 2.0

Table 4 2.0 g where in the Neilson study the abutment active action med-
Limit states [17]. ian PGA is 2.62 g.
Components Slight Moderate Extensive Complete From the component-level fragilities, the system level fragility
Elastomeric Bearing-Long (mm) 28.9 104.2 136.1 186.6
curve shown in Fig. 11(c) was constructed using Eq. (3). The system
Elastomeric Bearing-Tran (mm) 28.8 90.9 142.2 195 fragility curve for the current study is of the 15 bridges in the gen-
Abutment Trans/Long (mm) 9.8 37.9 77.2 NA erated bridge inventory. Similarly, Fig. 11(a) and 11(b) show the
single span bridge fragility of the Choi study [7] and the single span
bridge class of the Neilson study [19], respectively. The slight dam-
age state is the first damage level to be encountered by a bridge
past studies [6,17] using median PGA values. All component-level system, including bearing and abutment components. As seen in
median PGA values for the fixed bearing, expansion bearing, and Fig. 11(a), the slope of the slight damage fragility curve obtained
abutment from the current study, the Choi study, and the Neilson from the Choi [6] study is higher than those from the Nielson
study are summarized in Table 5. It should be noted that median and current study. This may be attributed to the fact that the Choi
PGA values above 4 g are considered insignificant because ground bridge had the longer span length, resulting in large seismic dis-
acceleration is unlikely to ever reach this level and are therefore placements at bearings in the longitudinal and transverse direc-
represented by the value 99 [17]. The Neilson study and the cur- tions. This indicates that slight damage will occur most rapidly
rent study incorporated several bridges into the single span class, over the 0.1–0.3 g range.
while the Choi [6] study only included a single PSC I-girder bridge. In Fig. 11(b) and (c), the slope of the slight damage curve is gen-
The most notable finding displayed in Table 5 is that the abutments tler, signifying a slower rate of damage proliferation. Likewise, the
for the current study show more fragility than in the other studies. three remaining damage states; moderate, extensive, and com-
The abutment has a media PGA for the slight damage state of 0.75 g plete, show similarity between the Nielson [17] study and the cur-
in the transverse direction where the Neilson study shows 1.38 g. rent study in terms of fragility curve slope and peak probability.
Likewise for slight damage, in the longitudinal direction for the For the slight damage state, the probability of exceedance at
abutment, the median PGA for the current study is on average 1.0 g is around 0.95 for the current study and 0.90 for the Neilson
J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 187

Fig. 10. Component fragility curves for: (a) fixed bearing; (b) expansion bearing; (c) abutment with fixed bearing; and (d) abutment with expansion bearing.

The median PGA values taken from the fragility curves of Fig. 11
Table 5 for each study are displayed in Table 6. In the current study, under
Median PGA values for bridge components. the complete damage state, the median is 2.75 g, whereas the value
Slight Moderate Extensive Complete is 2.50 g for the Neilson study. This is a percent difference of 9%.
When comparing the Neilson study with the current study, the
Component Median (g)
percent difference is 5% for the extensive damage state, 0.7% for
Choi [6] Fxd-Long 0.30 0.63 0.82 1.13
the moderate state, and 10% for the slight damage state. As shown
Fxd-Trans 0.14 0.61 0.77 1.06
Exp-Long 99 99 99 99 in Fig. 11 and Table 6, the results correlate well between the
bridges in the current study and the work of Neilson [17]. Consid-
Neilson [17] Fxd-Long 0.66 2.10 2.67 3.56
Fxd-Trans 99.00 99.00 99.00 99.00 ering that single span bridges are less fragile than multi-span
Exp-Long 0.62 2.11 2.72 3.68 bridges, [7,19] the effects of curvature are not as easily noticed in
Exp-Trans 1.46 99.00 99.00 99.00 a fragility analysis.
Ab-Pass 99.00 99.00 99.00 99.00
As seen in Table 6, median values for the bridge in the Choi [6]
Ab-Act 2.62 99.00 99.00 99.00
Ab-Tran 1.38 99.00 99.00 99.00
study shows extremely low values indicating a fragile bridge. The
percent differences between Choi and the current study are
Current Study Fxd-Long 1.40 99.00 99.00 99.00
Fxd-Trans 1.68 99.00 99.00 99.00
100%, 90%, 100%, and 104%, for the slight, moderate, extensive,
Exp-Long 1.31 99.00 99.00 99.00 and complete damage states, respectively. The results differ when
Exp-Trans 1.68 99.00 99.00 99.00 compared to the Choi study [7] mainly due to the differences in
Ab-Fxd-Long 1.70 99.00 99.00 99.00 bridge geometry that were outlined in Table 2a. In addition, the
Ab-Fxd-Trans 0.74 2.63 99.00 99.00
Choi study only looked into seismic fragility of a single PSC bridge,
Ab-Exp-Long 2.33 99.00 99.00 99.00
Ab-Exp-Trans 0.75 2.66 99.00 99.00 whereas the current study investigated the bridge portfolio’s
fragility.
Note: Fxd = Fixed Bearing; Exp = Expansion Bearing; Ab = Abutment; Long = Longi-
tudinal; Trans = Transverse.

7. Conclusions and future work

study. The Choi [6] study displays higher fragility than the other This paper proposed two modeling approaches, including the
two studies mentioned as indicated by the more aggressive slopes area and spine modeling methods, for seismic fragility analysis of
and greater peak probabilities of the fragility curves for the moder- curved precast-prestressed concrete (PSC) bridge portfolios and
ate, extensive and complete damage levels. compared the results. These were developed based upon finite ele-
188 J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189

efficiency between the two approaches along with the comparison


to past studies [6,17]. Component- and system-level fragility
curves of a portfolio of single span, curved, PSC I-girder bridges
modeled using the spine modeling approach were constructed.
These curves were compared to those gained from the past studies
to identify similarities and verify the modeling approach that can
be efficiently used for seismic vulnerability analysis for such bridge
portfolios. Findings for this study are relevant to the presented
models and modeling techniques. It is important to realize that
there are many finite element modeling methods available and
that these findings may not necessarily apply to all other methods.
Significant findings from this study are as follows:
Comparison between the two approaches showed that the spine
modeling approach was considered more efficient to model the
selected curved PSC bridges and simulate their seismic behaviors
than the area model. Specifically, all the critical component seismic
responses obtained from the two approaches in the transverse
direction were almost identical, though preliminary seismic results
indicated a somewhat discrepancy between displacements in the
vertical and longitudinal directions. However, it is the responsibil-
ity of the curved PSC bridge design engineer to determine which
modeling technique is appropriate for any given situation. Factors
affecting this decision can include computational efficiency, impor-
tance factor of the structure, seismic zone, level of detail desired,
and structure complexity.
The efficiency comparison determined that the spine model has
several advantages over the area model. Specifically, a total time
saving of 7% was shown when using the spine model for a curved
PSC bridge. This would be significant for seismic vulnerability stud-
ies for PSC bridge portfolios with vastly different configurations.
The component- and system-level fragility curves of the PSC I-
girder bridge portfolio were compared to those from the past stud-
ies. The fragility analysis turned out to be comparable among the
studies, and the comparison of these system fragilities solidifies
the spine modeling approach for regional curved PSC bridges as a
practical option.
A comprehensive parametric study to investigate the effect of
bridge characteristic parameters on seismic fragility of single-
span curved PSC bridges should be further conducted. Because ver-
tical seismic fragility of bridges has been in question for many
years, the future parametric study should include vertical fragility
analyses to provide evidence to support or deny that single-span
curved PSC bridges are vulnerable to vertical ground motions.
Meanwhile, the use of area modeling approach in a fragility analy-
sis considering the membrane effects and plate-bending behaviors
for the curved superstructure system could be further investigated
Fig. 11. System level fragility curve comparison for single span bridge class: (a)
Choi [6]; (b) Neilson [17]; and (c) current study. to determine any increases in accuracy. Because this would be
computationally intensive, appropriate experimental design tech-
niques are required. Finally, parameterized seismic fragility analy-
ment modeling techniques and were applied to a simply sup- sis enabling to account for uncertainty in configurational
ported, curved PSC I-girder bridge with radius of 402 m subjected parameters of curved PSC bridges should be investigated in an
to a suite of horizontal and vertical ground motions. The focus attempt to adequately assess their effects on structural
was to monitor changes in seismic responses and computational vulnerabilities.

Table 6
System-level median PGA values for different damage states.

Slight (g) %dif Moderate (g) %dif Extensive (g) %dif Complete (g) %dif
Choi [6] 0.13 100.00 0.49 89.89 0.63 100.40 0.87 103.87
Nielson [17] 0.35 10.81 1.30 0.77 1.80 5.41 2.50 9.52
Current Study 0.39 1.29 1.90 2.75

%dif = percent difference between past study and current study.


J. Seo, L.P. Rogers / Engineering Structures 150 (2017) 176–189 189

References [19] Nielson B, DesRoches R. Analytical fragility curves for typical highway bridge
classes in the Central and Southeastern United States. Earthquake Spectra
2007;23(3):615–33.
[1] AASHTO. LRFD bridge design specifications, Washington DC; 2012.
[20] Nielson BG, DesRoches R. Seismic fragility methodology for highway bridges
[2] Abdelnaby A, Raji F, Yohannes A, Naimi A, Mishra S, Golias M. Impacts of the
using a component level approach. Earthquake Eng Struct Dynam
1811–1812 earthquakes on existing transportation networks in Memphis area.
2007;36:823–39.
In: 10NCEE, tenth U.S. national conference on earthquake engineering frontiers
[21] Padgett J. Seismic vulnerability assessment of retrofitted bridges using
of earthquake engineering, 21–25 July, 2014 Anchorage, Alaska; 2014.
probabilistic methods [Ph.D. thesis]. Atlanta: Georgia Institute of
[3] Amirihormozaki E, Pekcan G, Itani A. Analytical fragility functions for
Technology; 2007.
horizontally curved steel I-girder highway bridges. In: Seventh national
[22] Padgett J, DesRoches R. Bridge functionality relationships for improved seismic
seismic conference on bridges and highways, 20–22 May, 2013, Oakland,
risk assessment of transportation networks. Earthquake Spectra
California; 2013.
2007;23:115–30.
[4] Caltrans. Caltrans seismic design criteria version 1.3, 1st ed. Sacramento, CA:
[23] Padgett J, DesRoches R. Methodology for the development of analytical
California Department of Transportation; 2004.
fragility curves for retrofitted bridges. Earthquake Eng Struct Dynam 2008;37
[5] Central US Earthquake Consortium (CUSEC). Earthquake vulnerability of
(8):1157–74.
transportation systems in the Central United States; 2000.
[24] Precast/Prestressed Concrete Institute (PCI). Bridge design manual. 3rd
[6] Choi E. Seismic analysis and retrofit of mid-america bridges [PhD
ed. Chicago, IL: First Release; 2011.
thesis]. Atlanta, Georgia: School of Civil and Environmental Engineering,
[25] Ramanathan K, DesRoches R, Padgett JE. Analytical fragility curves for
Georgia Institute of Technology; 2002.
multispan continuous steel girder bridges in moderate seismic zones. Transp
[7] Choi E, DesRoches R, Nielson B. Seismic fragility of typical bridges in moderate
Res Rec : J Transp Res Board 2010;2202:173–82.
seismic zones. Eng Struct 2004;26(2):187–99.
[26] Ramanathan K, DesRoches R, Padgett JE. A comparison of pre- and post-seismic
[8] Computers and Structures Inc., 2014. CSIBridge v16; 2014.
design considerations in moderate seismic zones through the fragility
[9] DesRoches R, Delemont M. Seismic retrofit of simply supported bridges using
assessment of multispan bridge classes. Eng Struct 2012;45:559–73.
shape memory alloys. Eng Struct 2002;24:325–32.
[27] Seo J. Seismic vulnerability assessment of a family of horizontally curved steel
[10] Engineering Seismology Laboratory. SUNY at Buffalo. Buffalo, New York; 2005.
bridges using response surface metamodels [Ph.D. thesis]. University Park,
<http://civil.eng.buffalo.edu/engseislab/index.htm> [accessed March 2014].
Pennsylvania: Department of Civil and Environmental Engineering, The
[11] Ghosh J, Padgett JE. Aging considerations in the development of time-
Pennsylvania State University; 2009.
dependent seismic fragility curves. J Struct Eng 2010;136(12):1497–511.
[28] Seo J. Statistical determination of significant curved I-girder bridge seismic
[12] Ghosh J, Rokneddin K, Padgett JE, Dueñas-Osorio L. Seismic reliability
response parameters. Earthquake Eng Eng Vibr 2013;12(2):251–60.
assessment of aging highway bridge networks with field instrumentation
[29] Seo J, Linzell D. Horizontally curved steel bridge seismic vulnerability
data and correlated failures. I: methodology. J Earthquake Spectra 2014;30
assessment. Eng Struct 2012;34:21–32.
(2):795–817.
[30] Seo J, Linzell D. Use of response surface metamodels to generate system level
[13] Han R, Li Y, van de Lindt J. Seismic risk of base isolated non-ductile reinforced
fragilities for existing curved steel bridges. Eng Struct 2013;52:642–53.
concrete buildings considering uncertainties and mainshock–aftershock
[31] Seo J, Linzell D. Nonlinear seismic response and parametric examination of
sequences. Struct Saf 2014;50:39–56.
horizontally curved steel bridges using 3D computational models. J Bridge Eng
[14] Jeon J-S, Shafieezadeh A, Lee DH, Choi E, DesRoches R. Damage assessment of
2013;18(3):220–31.
older highway bridges subjected to three-dimensional ground motions:
[32] Seo J, Park H. Probabilistic seismic restoration cost estimation for
characterization of shear-axial force interaction on seismic fragilities. Eng
transportation infrastructure portfolios with an emphasis on curved steel I-
Struct 2015;87:47–57.
girder bridges. Struct Saf 2017;65:27–34.
[15] Jeon J-S, DesRoches R, Kim Taesik, Choi E. Geometric parameters affecting
[33] Shinozuka M, Feng MariaQ, Kim HK, Kim SHM. Nonlinear static procedure for
seismic fragilities of curved multi-frame concrete box girder bridges with
fragility curve development. J Eng Mech 2000;126(12):1287–96.
integral abutments. Eng Struct 2016;122:121–43.
[34] Torres E, Seo J, Rogers L. Probabilistic structural performance evaluation of
[16] Minnesota Department of Transportation (MNDOT), LRFD bridge design
concrete slab bridge system subjected to scour and earthquake. ACI Spec Publ
manual. Minnesota Department of Transportation 3485 Hadley Avenue
2017;316:73–94.
North Mail Stop 610 Oakdale, MN 55128-3307; 2014.
[35] Wilson T, Mahmoud H, Chen S. Seismic performance of skewed and curved
[17] Nielson B. Analytical fragility curves for highway bridges in moderate seismic
reinforced concrete bridges in mountainous states. Eng Struct
zones [Ph.D. thesis]. Atlanta, Georgia: School of Civil and Environmental
2014;70:158–67.
Engineering, Georgia Institute of Technology; 2005.
[36] FEMA. MH MR4 technical manual: Multi-hazard loss estimation methodology
[18] Nielson B, DesRoches R. Seismic performance assessment of simply supported
earthquake model. Dept. of Homeland Security Emergency Preparedness and
and continuous multispan concrete girder highway bridges. J Bridge Eng
Response Directorate. Washington, DC: FEMA Mitigation Div.; 2003.
2007;2(5):611–20.

You might also like