Вы находитесь на странице: 1из 36

Bull Earthquake Eng (2011) 9:14631498

DOI 10.1007/s10518-011-9260-8
ORIGINAL RESEARCH PAPER

A performance-based adaptive methodology


for the seismic evaluation of multi-span simply
supported deck bridges
Donatello Cardone Giuseppe Perrone
Salvatore Sofia

Received: 18 December 2009 / Accepted: 22 March 2011 / Published online: 9 April 2011
Springer Science+Business Media B.V. 2011

Abstract A performance-based adaptive methodology for the seismic assessment of highway bridges is proposed. The proposed methodology is based on an Inverse (I), Adaptive
(A) application of the Capacity Spectrum Method (CSM), with the capacity curve of the
bridge derived through a Displacement-based Adaptive Pushover (DAP) analysis. For this
reason, the acronym IACSM is used to identify the proposed methodology. A number of
Performance Levels (PLs), for which the seismic vulnerability and seismic risk of the bridge
shall be evaluated, are identified. Each PL is associated to a number of Damage States (DSs)
of the critical members of the bridge (piers, abutments, joints and bearing devices). The
IACSM provides the earthquake intensity level (PGA) corresponding to the attainment of
the selected DSs, using over-damped elastic response spectra as demand curves. The seismic
vulnerability of the bridge is described by means of fragility curves, derived based on the
PGA values associated to each DS. The seismic risk of the bridge is evaluated as convolution
integral of the product between the fragility curves and the seismic hazard curve of the bridge
site. In this paper, the key aspects and basic assumptions of the proposed methodology are
presented first. The IACSM is then applied to nine existing simply supported deck bridges,
characterized by different types of piers and bearing devices. Finally, the IACSM predictions
are compared with the results of nonlinear response time-history analysis, carried out using
a set of seven ground motions scaled to the expected PGA values.
Keywords Bridges Seismic evaluation Damage states Adaptive pushover analysis
Capacity spectrum method Nonlinear response time-history analysis

1 Introduction
Recent earthquakes have repeatedly demonstrated the seismic vulnerability of existing
bridges, due to their design based on gravity loads only or inadequate levels of lateral forces

D. Cardone (B) G. Perrone S. Sofia


DiSGGUniversity of Basilicata, Potenza, Italy
e-mail: donatello.cardone@unibas.it

123

1464

Bull Earthquake Eng (2011) 9:14631498

(Priestley et al. 1996). The seismic zonation of many European countries, moreover, has
been revised recently, prescribing more severe Peak Ground Accelerations (PGAs) in several
regions. Reliable methods for the assessment of the vulnerability and seismic risk of existing
bridges, to be used within a maintenance and management framework of a road network, are
then needed.
It is well known that the most accurate method of analysis for the evaluation of the
seismic response of structures is the Nonlinear response Time-History Analysis (NTHA).
This technique, however, requires the preliminary selection of an appropriate set of ground
motions and considerable computational efforts with the consequent difficulty of producing
ready-to-use results. Nonlinear Static Procedures (NSPs) based on pushover analysis represent nowadays a reliable alternative to NTHA and are widely recognized as a practical, yet
accurate, engineering tool for the assessment of existing structures.
The general purpose of NSPs is the evaluation of the so-called Performance Point (PP),
defined, in seismic assessment terms, as the damage level of the structure due to a seismic
event of predefined seismic intensity. From a graphical point of view, the PP can be derived by
intersecting the capacity spectrum of the structure, derived from pushover analysis, with the
seismic demand of the expected ground motions, represented by suitable response spectra,
in the so-called ADRS (Acceleration Displacement Response Spectra) format.
The most widely used NSPs for the seismic assessment of structures are: the Capacity
Spectrum Method (CSM) (ATC 1996), the Displacement Coefficient Method (DCM) (ATC
2005) and the N2 method (Fajfar 1999; CEN 2005). All these methods are based on the
conversion of the pushover curve of the multi-degree-of-freedom (MDOF) nonlinear model
of the structure into an equivalent single-degree-of-freedom (SDOF) idealized (e.g. bilinear)
capacity curve, based on the sole first mode of the structure. It should be noted here that the
idea of representing a structure (such a multi-storey building or a multi-span bridge) by an
equivalent SDOF oscillator, whose characteristics are defined from a nonlinear static analysis
of the corresponding MDOF structure, was first explored by Saiidi and Sozen (1981), who
extended previous works by Biggs (1964) and Clough and Penzien (1975) on equivalent
SDOF systems.
The currently used NSPs basically differ in the bilinearization technique adopted and in
the description of the seismic demand by either high-damped elastic or isoductile inelastic response spectra. NSPs have been originally developed and applied to buildings. Only
recently the applicability of NSPs to bridges has been comprehensively examined (Paraskeva
et al. 2006; Casarotti et al. 2005; Aydinoglu and nem 2007).
Traditional pushover analysis techniques, based on monotonically increasing invariant
distributions of lateral load patterns are suitable for structures whose seismic response is
dominated by the first mode of vibration. In bridges, the effects of higher modes are generally not negligible and the modes of vibration of the structure can change significantly
during strong earthquakes, due to damage of structural members (piers, bearing devices,
abutments, etc.). Recent studies on the importance of higher mode effects and the accuracy
of NSPs based on single-mode pushover analysis (such as the N2 method adopted in the
European seismic code) have been carried out by Isakovic et al. (2003, 2008), Isakovic and
Fischinger (2006), on a number of continuous single column bent viaducts. The results of
these studies clearly point out that the higher modes significantly influence the response
of relatively long viaducts, regardless the seismic intensity and the structural characteristics of the bridge (stiffness ratio between deck and bent, eccentricity between centre of
mass and stiffness, ratio between torsional and translational stiffness, type of constraints
at the abutments), thus making the use of NSPs based on single-mode pushover analysis
inadequate.

123

Bull Earthquake Eng (2011) 9:14631498

1465

Different pushover analysis techniques, based on multimodal and/or adaptive approaches,


are then needed, to account for the higher modes effects and the redistribution of inertia forces
caused by structural damage and the associated changes in the vibration properties of the
structure (Kappos and Paraskeva 2008).
In the modal pushover analysis (MPA) proposed by Chopra and Goel (2002), a number
of simultaneous pushover analyses are carried out separately for each significant mode. The
contributions from individual modes to the seismic response of the structure are combined
using appropriate combination rules (SRSS or CQC). Recently, Paraskeva and Kappos (2010)
proposed some changes to MPA, based on the use of the so-called modal inelastic deformed
shape of the structure, in lieu of the elastic modal shape, to adapt the use of MPA to the
seismic assessment of bridges.
Aydinoglu (2003), Aydinoglu and nem (2007) proposed the so-called Incremental
Response Spectrum Analysis (IRSA), wherein, each time a significant change of stiffness
occurs in the structure, an elastic modal spectrum analysis is performed, taking into account
the changes in the dynamic properties of the structure.
First adaptive approaches have been presented by Reinhorn (1997) and Elnashai (2001).
More recently, an innovative Displacement-based Adaptive Pushover (DAP) technique
has been proposed by Antoniou and Pinho (2004a,b), in which a set of lateral displacements (rather than forces) is monotonically applied to the structure. The displacement pattern is updated at each step of the analysis, based on the current dynamic
characteristics of the structure. Pinho et al. (2007) demonstrated that the use of the DAP
technique can lead to the attainment of significantly improved predictions, compared to
conventional (i.e. force-based) adaptive pushover procedures, which match very closely
results from nonlinear response time-history analysis, even and especially for irregular
bridges.
An adaptive re-interpretation of the CSM has been also recently proposed by Casarotti
and Pinho (2007) for bridge applications. The proposed method, called ACSM (Adaptive
Capacity Spectrum Method), features two types of adaptiveness. First, the pushover algorithm. Indeed, the DAP analysis is adopted to better estimate the inelastic bridge behaviour, independently of structural regularity. The second element of adaptiveness of the
ACSM resides in the way the pushover curve is converted in the equivalent SDOF idealized capacity curve. Indeed, all the equivalent SDOF quantities vary based on the current
deformed shape of the bridge, in accordance with the principles of the Direct Displacement-based Design (DDBD) method (Priestley et al. 2007). In the traditional CSM, on the
contrary, the conversion is carried out based on the sole elastic modal properties of the
structure.
In this paper, a performance-based adaptive methodology for the evaluation of the seismic vulnerability and seismic risk of highway bridges is proposed. In the current version, the proposed methodology is specialised to multi-span bridges, with either simple
(isostatic) or continuous (hyperstatic) deck supported on bearings, although it could be
extended to other bridge types simply by revising some modelling assumptions, as discussed
below.
In the first part of the paper, the background and implementation of the proposed methodology are presented. In the second part of the paper, the proposed methodology is applied to
a set of nine simply supported deck bridges of the Italian A16 NapoliCanosa highway, characterised by different types of piers and bearing devices. The effectiveness of the proposed
methodology in reproducing the global behaviour and local phenomena of existing bridges
is assessed by comparing the expected bridge response with the results of accurate nonlinear
response time-history analyses.

123

1466

Bull Earthquake Eng (2011) 9:14631498

2 Description of the proposed methodology


2.1 Key aspects of the proposed methodology
The proposed methodology is based on an Inverse (I) application of the ACSM. For this
reason, the acronym IACSM is used to identify the proposed methodology. Contrary to
ACSM, however, the IACSM is not iterative and does not require the bilinearization of the
capacity curve of the equivalent SDOF model of the bridge. In the direct application of the
ACSM, indeed, the main objective is to find the performance point (i.e. the damage state) of
the structure under a predefined seismic intensity (i.e. a given PGA) using an over-damped
demand spectrum derived by an iterative procedure on an idealized bi-linear capacity curve
(ATC 1996). The main objective of the IACSM, instead, is to evaluate the seismic intensity
(i.e. the PGA) of the expected ground motions, corresponding to pre-determined damage
states of the structure, identified by given performance points on the capacity curve of the
bridge. The inelastic deformed shape of the bridge corresponding to each damage state, therefore, is already known at the beginning of the analysis. As a consequence, the equivalent
damping ratio of the bridge can be directly evaluated by properly combining (see Sect. 2.3.)
the damping contributions of the single bridge components.
In the proposed methodology, two Displacement Adaptive Pushover (DAP) analyses are
performed to derive the capacity curves of the bridge in the longitudinal and transverse
direction, respectively. The DAP technique has been preferred to the other conventional
(i.e. force-based) adaptive pushover techniques, to better estimate the inelastic deformed
shape of the bridge and the distribution of the shear forces between the structural elements.
Compared to multiple-run pushover techniques (such as the MPA and IRSA methods), the
single-run DAP technique results decidedly more suitable for an inverse application of the
CSM. Moreover, it does not require the definition of any reference point.
Basically, the DAP algorithm consists of four steps: (i) definition of a nominal displacement vector (Ui,0 , i = 1, , n), (ii) computation of a load increment factor (k ), (iii)
calculation of a normalized modal scaling vector (Di,k , i = 1,, n) and (iv) update of the
displacement vector (Ui,k , i = 1, , n), where Ui and Di represent the absolute displacements
of the center of mass of the ith span of the bridge and n stands for the total number of spans.
Whilst the first stage is carried out only once, at the start of the analysis, its three remaining
counterparts are repeated at every (kth) step of the nonlinear static analysis procedure.
A uniform deformed shape of the bridge has been assumed as nominal displacement vector. The normalized modal scaling vector Di,k , which defines the shape (not the magnitude)
of the displacement increment vector, is computed at the start of every load increment as:
Di,k =

Di,k


max Di,k

(1)

In order for such modal scaling vector to reflect the actual stiffness state of the structure, as
computed at the end of the previous load increment, an eigenvalue analysis is first carried out
to determine the modal shapes (ij ) and modal participations factors (j ) of the structure.
Modal results are combined using the SRSS combination rule:




 m 2
 m 
2
Di,k = 
j,k ij,k Sd (Tj,k )
Dij,k = 
(2)
j=1

j=1

where Sd (Tj,k ) represents the displacement response spectrum ordinate corresponding to the
period of vibration of the jth mode and m stands for the total number of modes.

123

Bull Earthquake Eng (2011) 9:14631498

1467

The displacement vector at a given step k of the DAP analysis is updated according to an
incremental response control algorithm:
Ui,k = Ui,k1 + k Di,k Ui,0

(3)

until a predefined analysis target or numerical failure is reached.


The DAP algorithm has been implemented in SeismoStruct (SeismoSoft 2006), which
can be freely downloaded from the Internet. Full details on this computer package can be
found in its accompanying manual (SeismoSoft 2006).
One of the key aspects of the proposed methodology is the definition of a number of
Performance Levels (PLs), for which the vulnerability and seismic risk of the bridge shall
be evaluated. Each PL is associated to a number of Damage States (DSs) of the critical
members of the bridge, identified by a series of points on the DAP curve of the bridge. The
IACSM provides the earthquake intensity levels (PGA) corresponding to the attainment of
the selected DSs, using over-damped elastic response spectra as demand curves. The seismic
vulnerability of the bridge is then described by means of fragility curves, derived based on
the PGA values previously obtained for each DS. Finally, a seismic risk index is evaluated
as convolution integral of the product between the fragility curves and the seismic hazard
curve of the bridge site.
As far as the bridge modelling is concerned, the so-called Structural Components Modelling (SCM) approach (Priestley et al. 1996) has been followed, in which a number of idealized
structural subsystems are connected to resemble the general geometry of the bridge. The phenomenological response of each structural subsystem is provided in the form of member end
force-deformation relationships.
2.2 Bridge modelling assumptions
According to the SCM approach, the bridge structure can be divided in a number of independent rigid diaphragms, modelling the bridge decks, mutually connected by means of a
series of nonlinear springs, modelling bearing devices, piers and abutments (see Fig. 1). The
translational and rotational mass of each deck (equal to L and L3 /12, respectively,
where L is the total length of the deck and the seismic mass per unit of length) are lumped
in the centre of mass of each deck. If the total mass of the pier is more than 1/5 of the mass
of the part of the deck carried by that pier, a tributary mass of the pier (equal to the sum of
the mass of the pier cap and one third of the mass of the pier shaft) can be taken into account.
Table 1 summarises the basic modelling assumptions adopted for each bridge component, to describe their monotonic and cyclic behaviour, within nonlinear static and nonlinear
dynamic analysis, respectively.
As can be seen in Table 1, decks and foundations have been considered infinitely rigid and
resistant. The hypothesis of high seismic resistance of decks and foundations is related to their
low seismic vulnerability, compared to piers, abutments and bearing devices. The hypothesis
of high elastic stiffness, on the other hand, is substantiated by current seismic codes [e.g. see
EC8-part 2 (CEN 2004)], which gives the possibility to assume a rigid deck model when the
deformation of the deck in a horizontal plane is negligible compared to the displacements
of the pier tops. This is always valid in the longitudinal direction of approximately straight
bridges. In the transverse direction, the deck can be assumed as rigid, according to EC8, if
L/B < 4 or in general if Dd /da < 0.2, where L and B are the length and the width of the deck,
respectively, Dd and da are the maximum difference and the average of the displacements in
the transverse direction of all the pier tops under the transverse seismic action. It is worth
to note that the aforesaid conditions for the applicability of the rigid deck model are always

123

1468

Bull Earthquake Eng (2011) 9:14631498

(a)
Pushing action

Pushing action
Deck mass

Bearing

Pier mass

Bearing

Pier mass
Pier

ag

Joint

Deck mass

Pulling action

Pier

(b)

Bearing

Bearing
Pulling action

Deck mass

Joint

Joint

L
L2 /12

Joint

Joint

Bearing

Bearing

Bearing

Deck mass

ag

Pier mass

L
L2 /12

Joint

Bearing

Pier mass
Pier

Pier

Fig. 1 Schematization of the bridge structure for seismic analysis in the a longitudinal and b transverse
direction

F
Joint
Deck 1

Deck 2
Gap
Pier

Bearing

Fig. 2 Modelling of joints

satisfied, with adequate margin, for the nine simply supported deck bridges examined in this
study, which are representative of the Italian bridge inventory (ASPI 1992). Therefore, it can
be concluded that the hypothesis of infinitely rigid deck is, in most of the cases, appropriate
for simply supported deck bridges. In first approximation, also the hypothesis of infinitely
rigid foundation is reasonable in many cases. Again, this is substantiated by current seismic
codes [see EC8-part2 (CEN 2004)], which state that, generally, pier and abutments shall be
assumed as fixed to the foundation soil. Soil-structure interaction effects should be used only
when the displacement due to soil flexibility is greater than 30% of the total displacement at
the center of mass of the deck. The hypothesis of infinitely rigid and resistant foundation may
be removed in the next versions of the proposed methodology, by introducing appropriately
defined soil springs.
Possible effects due to the closure of the joints have been taken into account in the analyses. Indeed, gap closure results in pounding between adjacent decks and/or between deck
and abutment, which can determine a significant redistribution of the seismic forces between
piers and abutments. Joints have been modelled with compression-only translational (longitudinal direction) and rotational (transverse direction) link elements, with an initial gap
assigned based on the clearance of the joint (see Fig. 2). In the application of the proposed
procedure (see Sect. 3), a gap of 2050 mm has been considered, based on the actual width
of the expansion joints of the bridges examined.
Seat-type abutments founded on piles have been considered, since they represent the abutment type most widely used in Italy. The seismic response of the abutments in the longitudinal

123

Infinitely rigid and resistant


Lumped translational and rotational mass

Compression-only translational (long. dir.)


and rotational (transv. dir.) springs with gap

Transv. dir.: infinitely rigid and resistant

Deck

Joints

Abutments

Bearings

Piers

Infinitely rigid and resistant

Foundations

Post-failure pier-deck sliding considered

Flexural behaviour derived from


moment-curvature analysis
Shear strength and P- effects considered

Beam with plastic hinges at the end(s)

Long. dir.: effects of piles-ground and


backfill-abutment interaction considered

Modelling assumptions

Component

Viscous, hysteretic or frictional cyclic


behaviour, depending on bearing type and
displacement amplitude

Takeda degrading-stiffness hysteretic model

Long. dir.: kinematic multilinear plastic


model

Rigid behaviour after gap closure

Elastic-plastic with hardening backbone


curve

Multilinear backbone curve

Cyclic behaviour (NTHA)

Diaphragm behaviour

Long. dir.: elastic-perfectly-plastic backbone


curve

Monotonic behaviour (NSA)

Table 1 Basic modelling assumptions considered in the nonlinear static analysis (NSA) and nonlinear time-history analysis (NTHA) of bridges

Bull Earthquake Eng (2011) 9:14631498


1469

123

1470

Bull Earthquake Eng (2011) 9:14631498

Deck
Abutment

Back Fill

Back Wall

Bearing

Piles

Pushing action

Pulling action

Fig. 3 Mechanical behaviour of abutment under pushing and pulling action

direction has been captured with a couple of nonlinear springs, characterised by two different elastic-perfectly-plastic backbone curves (see Fig. 3), modelling the pushing and pulling
action of the abutment, respectively. The longitudinal response of the abutment is based on
the interaction between bearing devices, joint gap, abutment back wall, abutment piles and
soil backfill material. Prior to gap closure, the deck force is transmitted through the bearing
devices to the abutment wall and subsequently to piles and backfill, in a series system. After
gap closure, the deck pushes directly on the abutment back wall, with the risk of mobilizing the passive backfill pressure. In this study, the horizontal stiffness and ultimate strength
of the abutment in the pushing direction have been derived from a combination of design
recommendations (Caltrans 2006) and experimental test results on seat-type abutments with
piles (Maroney et al. 1993). They are expressed as a function of the abutment back wall
dimensions (w: width, hw : height) as follows:


hw
Kpush = ki w
(4)
1.7


hw
(5)
Rpush = hw w pi
1.7
in which ki and pi are two coefficients assumed equal to 11.5 (KN/ mm)/m and 239 kPa,
respectively, according to the Caltrans (2006) recommendations.
The horizontal stiffness of the abutment-pile system in the pulling direction has been
assumed equal to 7np KN/mm, np being the number of piles, in accordance with (DesRoches
et al. 2003). The horizontal strength of the abutment-pile system in the pulling direction has
been taken equal to the ultimate resistance of the pile group.
Piers have been modelled with nonlinear springs characterised by a bilinear backbone
curve. In this study, the lateral force-displacement relationships of the piers have been derived
based on preliminary elasto-plastic pushover analyses of the piers, schematized as elastic
beams with plastic hinges at the end(s). In the pushover analysis, the top displacement of
each pier is progressively increased, tracking the formation of plastic hinges up to their ultimate rotation capacity or the attainment of the pier shear strength. P  effects due to
gravity loads are considered in the analysis. First, a moment-curvature analysis of the critical
cross sections of the pier is performed, considering the axial force due to gravity loads and
the effects of concrete confinement and steel strain-hardening. In this study, reference to
the model by Mander et al. (1988) and Menegotto and Pinto (1973) has been made for confined/cover concrete and steel, respectively. The moment-curvature relationship thus obtained
is properly bilinearized (see Fig. 4a) and then converted in the moment-rotation behaviour of

123

Bull Earthquake Eng (2011) 9:14631498

(a)

1471

(b)

(c)

rK

Fy

Mu
My

High shear resistance

Kp2

Flexural behaviour
with P- effects

Mr

Kp3

Buckling and
lap-splice effects

Kp1
Low shear resistance

Fig. 4 Pier modelling: a moment-curvature analysis of the critical section, b force-displacement behaviour
taking into account possible shear failure, c Takeda degrading-stiffness-hysteretic model

the plastic hinge, whose length (Lp ) has been evaluated according to the formula by Priestley
et al. (1996):
Lp = 0.08L + 0.022fyh dbl 0.044fyh dbl

(6)

where L is the distance between the centre of the plastic hinge and the first counterflexure
point in the elastic deformed shape of the pier, fyh and dbl are the yielding stress and the
minimum diameter of the steel longitudinal reinforcement, respectively.
In this phase, possible premature failure due to lap splice effects or buckling of longitudinal bars can be considered (see Fig. 4a) through simple semi-empirical relationships
(Priestley et al. 1996). After that, the shear resistance of the pier (Vd ) is computed, based on
the relationship proposed by Priestley et al. (2007):
Vd = Vc + Vs + Vp

(7)

in which Vc is the shear resistance of concrete, Vs the contribution of the transverse reinforcement and Vp is the shear absorbed by the axial load. More details on the three shear
components (Vc , Vs and Vp ) can be found in Priestley et al. (2007).
Finally, the shear resistance, expressed as a function of the pier top displacement, is compared to the flexural behavior of the pier derived from elasto-plastic pushover analysis, in
order to determine the actual flexural/shear behavior of the pier (see Fig. 4b).
As far as the cyclic behaviour of the piers is concerned, reference to the Takeda degradingstiffness-hysteretic model has been made (see Fig. 4c). The Takeda degrading-stiffness-hysteretic model is very similar to the multi-linear Takeda model (Takeda et al. 1970) but has
additional parameters to control the degrading hysteretic loop. It is particularly well suited
for reinforced concrete members and it is based on the observation that unloading and reverse
loading tend to be directed toward specific points, called pivot points, in the force-deformation
plane. The Takeda degrading-stiffness-hysteretic model does take into account the strength
degradation due to P-delta effects (not shown in Fig. 4c), while it neglects the strength degradation due to cyclic effects near collapse. The Takeda degrading-stiffness-hysteretic model
is fully de scribed in (Dowell et al. 1998).
The nonlinear behaviour of the bearing devices is defined according to the bearing type
under consideration (see Fig. 5). If necessary, a post-failure frictional behaviour, corresponding to sliding between deck and pier cap, is considered.
Based on a comprehensive survey of the Italian highway bridge stock (ASPI 1992), which
comprises many bridges realised between 1960s and 1970s, five different types of bearing

123

1472

Bull Earthquake Eng (2011) 9:14631498

Fu
Ffr

Ffr

(a)

(b)

Ffr
K,

(c)

(d)

Fig. 5 Force-displacement behaviour of different types of bearing devices: a steel hinges and dowel steel
bars, b sliding bearings, c steel Pendulum and roller bearings and d neoprene pads

devices have been identified, i.e.: (i) steel hinges, (ii) dowel steel bars, (iii) sliding bearings,
(iv) steel pendulum and roller bearings, (v) neoprene pads.
Steel hinges and dowel steel bars are assumed to remain linear elastic up to failure (Fig. 5a),
which is usually brittle, being due to the attainment of the shear strength (Fu ) of the device.
The shear stiffness has been estimated based on the geometric details available. During the
analysis, the maximum shear force has been monitored. When the shear strength was prematurely exceeded, a post-failure frictional behaviour, corresponding to sliding between deck
and pier cap, has been considered (Fig. 5a).
Reference to a Coulomb (rigid-perfectly-plastic) model has been made to describe the
frictional behaviour of sliding bearings, steel pendulum and roller bearings (Fig. 5b, c).
As known, the friction coefficient (fr ) depends on many parameters and factors, such as:
bearing type, sliding interfaces, state of lubrication and maintenance, bearing pressure, air
temperature, sliding velocity, etc. (Dolce et al. 2005). As a consequence, a proper choice
of the friction coefficient shall be made, based on the data available (design specifications,
code requirements, manufacturer datasheets, common practice, etc.). In the application of
the proposed procedure to the selected case studies (see Sect. 3), a friction coefficient of
10% has been assumed for sliding and roller bearings (Fig. 5b), while it has been neglected
for steel pendulum bearings (Fig. 5c). Moreover, based on the geometric characteristics of
the examined bearing devices, the maximum displacement capacity of sliding/roller bearings
and steel pendulum bearings has been assumed equal to 150 and 100 mm, respectively.
A linear visco-elastic behaviour has been considered for neoprene pads (Fig. 5d), whose
shear stiffness (K) is evaluated based on the dimensions (cross section area and thickness) of
the pads and shear modulus (G) of neoprene. In this study, a shear modulus of 1 MPa and a
viscous damping ratio ( ) of 6% have been assumed as typical values of neoprene pads and
then adopted in the examples of application presented in Sect. 3. The horizontal strength of
the bearing system has been computed as the lowest between the shear resistance of neoprene

123

Bull Earthquake Eng (2011) 9:14631498

1473

Pushover analysis of the bridge

I. Define a number of Damage States (DSs)


II. Convert the Pushover Curve
in the Adaptive Capacity Spectrum

II.a Actual deformed shape of the


bridge at each (k-th) step of analysis
III.a Equivalent Damping PP
III.b Reduction Factor (PP)

III. Get the Demand Spectrum (for each DS)


IV.a Effective Period T PP

IV. Evaluate PGA (for each DS)

IV.b Target Acceleration Sa,PP


IV.c Spectral Acceleration Sa1(TPP)

V. Generate Fragility Curves (for each DS)

V.a Lognormal Cumulative


Probability Function

VI. Evaluate Seismic Risk (for each DS)


VI.a Seismic Hazard Curve

Fig. 6 Flowchart of the proposed methodology for the evaluation of the vulnerability and seismic risk of
bridge structures

pads and the friction resistance between neoprene and concrete sliding surfaces. In the case
studies examined in Sect. 3, the shear resistance of neoprene pads has been associated to the
attainment of a shear strain of 150%. The friction coefficient between neoprene and concrete
has been taken equal to 70%.
2.3 Evaluation of bridge vulnerability and seismic risk
The first step of the proposed procedure (see Fig. 6) is to define a number of performance
levels, for which the vulnerability and seismic risk of the bridge shall be evaluated. Each
PL is associated to a number of Damage States (DSs) of the critical members of the bridge
(piers, abutments and bearing devices), identified by a series of Performance Point (PP) on
the DAP curve of the bridge.
Herein, the PLs are divided in three groups (see Table 2), based on the consequences in
terms of damage that the attainment of each PL can produce. Obviously, this division is only
formal and it is made only for more clarity. Any point of the DAP curve of the bridge can be
arbitrarily selected at the beginning of the analysis.
The first group (PL1) includes post-earthquake Damage States (D1j in Table 2) in which
only very limited structural damage has occurred and although some minor structural repairs
may be appropriate, these generally do not require any traffic interruptions. Examples of DSs
of the first group are: (i) pier yielding, (ii) attainment of the horizontal strength in neoprene
pads (iii) closure of the joints, etc.
The second group (PL2) includes post-earthquake Damage States (D2j in Table 2) in which
significant damage to some structural elements has occurred but large margin against either
partial or global collapse still remains. The overall risk of life-threatening injury as a result
of structural damage is expected to be low. Whilst the damaged structure is not an imminent
collapse risk, it would be prudent to implement shortly structural repairs. This may require
traffic interruptions or the installation of temporary bracing systems. Examples of DSs of
the second group are: (i) attainment of the ultimate displacement capacity of sliding/roller

123

1474

Bull Earthquake Eng (2011) 9:14631498

Table 2 Selected performance levels (PLs) and corresponding damage states (DSs)
PL1 (slight damage
state)

PL2 (moderate
damage state)

PL3 (severe damage


state)

Joint closure (DS11)

Pier 50% ultimate ductility (DS21)

Failure of fixed hinges (DS31)

Pier yielding (DS12)

Displacement capacity of
sliding/roller bearings
(DS22)
Active resistance of abutment
(DS23)

Pier ultimate ductility (DS32)

Maximum force of neoprene


pads (DS13)

Pier shear failure (DS33)


Passive resistance of
abutment (DS34)
Deck unseating (DS35)

bearings, which is generally associated to the occurrence of large residual displacements


of the deck, (ii) significant ductility demands (e.g. 50% of the ultimate ductility) in piers,
(iii) attainment of the active resistance of abutments (pulling action), etc.
The third group (PL3) includes post-earthquake Damage States (D3j in Table 2) in which
the structure continues to support gravity loads but retains no margin against collapse. Aftershock activity could induce the partial or total collapse of the bridge. Extensive damage to
the structure has occurred, that implies significant degradation in the stiffness and strength of
the lateral-force resisting system and large permanent lateral deformations of the structure.
Significant risk of life-threatening injury exists. The structure may be technically repairable
but costs could be very high and the closure of the bridge for long time is inevitable. Examples
of DSs of the third group are: (i) collapse of steel hinges or dowel bar connections, (ii) pier
collapse due to the attainment of its ultimate ductility or shear strength, (iii) attainment of
the passive resistance of abutments (pushing action), (iv) incipient unseating of the deck, etc.
The second step of the procedure is to convert the pushover curve of the nonlinear MDOF
model of the bridge into an equivalent SDOF adaptive capacity spectrum. The latter is stepby-step derived by calculating the displacement (Sd,k ) and the acceleration (Sa,k ) of the
equivalent SDOF system based on the actual deformed shape of the bridge at each (k-th) step
of analysis, according to the fundamental equations of the DDBD method (Priestley et al.
2007) particularized to the bridge model considered:


2 + I 2
m

D
j
j
j
j,k
j,k

Sd,k =
(8)
m

D
j,k
j j
Sa,k =

Vb,k
Me,k g

(9)

where Vb,k is the base shear of the bridge, mj and Dj,k are the translational mass ( L)
and the horizontal displacement of the centre of mass of the jth deck, respectively, Ij and j,k
are the rotational mass ( L 3 /12) and the rotation around the vertical axis of the jth deck,
respectively, and Me,k is the effective mass of the entire bridge:

j mj Dj,k
(10)
Me,k =
Sd,k
The third step of the procedure is to determine the seismic demand associated to each DS,
represented by a reference over-damped elastic response spectrum, whose seismic intensity
(PGAPP ) is still unknown at this step of the analysis. This step requires the evaluation of the

123

Bull Earthquake Eng (2011) 9:14631498

1475

equivalent viscous damping of the bridge associated to each DS. To this end, the following
routine is followed: (i) find out the PP of the pushover curve corresponding to the selected DS,
(ii) enter the pushover database to determine the corresponding deformed shape of the bridge
and the actual displacement of each structural member, (iii) evaluate the equivalent damping
of each structural member, (iv) combine the damping contributions of each structural member
to get the equivalent viscous damping of the entire bridge.
The equivalent damping of the bearing devices can be calculated through the well-known
Jacobsen approach (Jacobsen 1930):
b,j (PP) =

Evisc + Ehyst + Efr


2 Fb,j db,j

(11)

in which Evisc , Ehyst and Efr identify the energy dissipated by the device, through its viscous,
hysteretic or frictional behaviour, in a cycle of amplitude db,j , being db,j the displacement of
the device at the selected DS and Fb,j the corresponding force level.
As far as piers are concerned, reference has been made to the following relationship:


1
(1 r )
p, j (PP) =
1
r
(12)

which relates the equivalent hysteretic damping of the pier to its displacement ductility ()
and post-yield hardening ratio (r). The aforesaid relationship has been derived by Kowalsky
et al. (1995), by applying the Jacobsens approach to the Takeda degrading-stiffness-hysteretic model.
The equivalent damping of each pier-bearings system is then computed, by combining the
damping values of pier and bearing devices in proportion to their individual displacements:
j (PP) =

b,j db,j + p,j dp,j


db,j + dp,j

(13)

Finally, the equivalent damping ratios of the pier-bearings systems are combined to provide
the equivalent damping ratio of the bridge as a whole, for the selected PP. The approach
followed is to weigh the damping values of the single pier-bearings systems in proportion to
the corresponding force levels (Fj = Fb,j = Fp,j neglecting pier mass):

n
j=1 j Fj
j=1 j Fj
PP =
n
=
(14)
Vb
j=1 Fj
Once the equivalent damping of the bridge has been determined, the corresponding demand
spectrum can be derived from the reference 5%-damping normalized response spectrum,
using a proper damping reduction factor (Cardone et al. 2008). As already suggested in previous studies on the seismic design of bridges (Kowalsky et al. 1995; Kowalsky 2002), in the
proposed procedure the damping reduction factor adopted in an old version of the Eurocode
8 (CEN 1998):

7
=
(15)
(2 + P P )
has been employed, with the limitation > 0.50, corresponding to a limit damping ratio of
approximately 25%.
The fourth step of the procedure is to determine the PGA values associated to each DS.
From a graphical point of view, this can be done by a translation of the over-damped Normalised Response Spectrum (NRS in Fig. 7a) to intercept the Adaptive Capacity Spectrum

123

1476

Bull Earthquake Eng (2011) 9:14631498

(a)
Sa (g)
Sa1 (TPP,PP)

P(DDS|PGA)

High-damping NRS

ACS

Sa,PP
PGAPP

(b)

5%-damping NRS

0.5

PP

PP

Demand
spectrum
Sd,PP

PGAPP

Sd

PGA

Fig. 7 a Evaluation of the PGA associated to a selected performance point (PP) and b derivation of the
corresponding fragility curve

(ACS in Fig. 7a) in the Performance Point (PP). From an analytical point of view, the PGA
associated to a given PP can be determined as the ratio between the acceleration of the ACS
corresponding to that PP (Sa,PP in Fig. 7a) and the normalized spectral acceleration at the
effective period of vibration (TPP ) and global equivalent damping (PP ) associated to the
selected PP (Sa1 (TPP , PP ) in Fig. 7a):
PGAPP =

Sa,PP
Sa1 (TPP , PP )

(16)

being:

TPP = 2

M
= 2
K


Sd,PP
g Sa,PP

(17)

The PGA values thus obtained represent an estimate of the median threshold value of the
peak ground acceleration related to the selected PP. They can be used to derive a number of
fragility curves (see Fig. 8b), which provide the probability of exceedance of the selected
DS, as a function of the PGA of the expected ground motions. In the proposed procedure,
the fragility curves are expressed by a lognormal cumulative probability function:



1
PGA
P(D DS |PGA ) = 
(18)
ln
c
PGAPP
in which P() is the probability of the Damage (D) being equal to or exceeding the selected
DS, for a given seismic intensity (PGA),  is the standard lognormal cumulative probability function, PGAPP the median threshold value of the seismic intensity associated to the
selected PP and c the lognormal standard deviation which takes into account the uncertainties related to input ground motion, bridge response, material characteristics, damage state
definition, etc. According to previous studies (Dutta and Mander 1998; Basz and Mander
1999; Kappos and Paraskeva 2008; Paraskeva and Kappos 2010), a value of c equal to 0.6
can be assumed for existing RC bridges. Lower values of c can be adopted for new bridges
or, generally speaking, when the knowledge of the structural characteristics is accurate.
The final step of the proposed procedure is the evaluation of the seismic risk associated
to the selected DS, with the use of hazard maps, which provide the PGA values at the bridge
site having a given probability of exceedance (e.g. 10%) in a given interval of time (e.g.
50 years). The seismic risk is obtained as convolution integral of the product between the

123

Bull Earthquake Eng (2011) 9:14631498

1477

(a)

(b)

Sa (g)

Vulnerability (V)
1
PP2

PGA3

PP1

PP3

PP2 PP3

0.5

PGA2
PGA1

PP1

PGA1 PGA2 PGA3

(c)

(d)

Hazard (P)

PGA

Ranking
R>R*

PxV

10%

PGA
PGA10%/50y

R<R*

R(PPi)

PGA

Fig. 8 Basic steps of the proposed procedure: a evaluation of the PGA related to given PLs, b derivation of
the corresponding fragility curves, c hazard curve of the bridge site and d calculation of the seismic risk index

seismic vulnerability of the bridge (V), expressed by the fragility curves (see Fig. 8b), and
the seismic hazard of the bridge site (P), expressed by a suitable hazard curve (see Fig. 8c):

R = P V dPGA
(19)
The risk index (R) thus obtained provides the probability of exceedance of the selected DS,
conditioned to the hazard of the bridge site. It can be used by the network manager to take
decisions concerning possible repair or seismic retrofit measures (see Fig. 8d).

3 Validation of the proposed methodology


3.1 Case studies
A set of 9 existing bridges of the A16 NapoliCanosa Italian highway has been selected as
case study. This selection resulted from a preliminary examination of the entire bridge inventory, aimed at identifying bridge types and bridge characteristics representative of the A16
Highway. It is worth to observe that approximately 90% of the Italian highway bridges is represented by multi-span simply supported deck bridges (ASPI 1992), either with independent
spans (by internal joints) or with continuous (cast in situ) top slab above the piers.
Figure 9 shows the schematic layout of the examined bridges. For each of them, bridge
geometry and pier type (SS: single shaft, SF: simple frame, SW: single wall) are outlined. In
Table 3 the main characteristics of each bridge structure are summarised. They include: (i)
pier type, (ii) bearing device type, (iii) span length, (iv) variability of pier heights (VH), (iv)
pier longitudinal and transverse reinforcement ratios (l and s respectively), (v) concrete

123

1478
Fiumarella

Bull Earthquake Eng (2011) 9:14631498

Hp,min=3.2m; Hp,max=3.2m; Ls=33.5m

SF
San Gennaro

Hp,min=10m; Hp,max=10m; Ls=33.5m

SW

Castello

Hp,min=6.3m; Hp,max=6.3m; Ls=33.5m

SW

Ceraso

Lauretta

Hp,min=4.6m; Hp,max=10m; Ls=33.5m

Hp,min=7m; Hp,max=11.8m; Ls=33.5m

SS

Macchione

Hp,min=3m; Hp,max=9m; Ls=33.5m

Serra dei Lupi

Hp,min=4m; Hp,max=5m; Ls=33.5m

SF

SF

Carapelle

Hp,min=4.5m; Hp,max=11.6m; Ls=33.5m

SW

Vallone delle Volpi

Hp,min=4.8m; Hp,max=4.8m; Ls=33.5m

SW

Fig. 9 Schematic layouts of the examined bridges

strength (fc ) and (vi) steel yielding (f y ). The variability of pier heights (VH) is evaluated
as percent difference between the average and minimum height of the piers of each span.
The highest value of VH, among all the spans of the bridge, is reported in Table 3. The
longitudinal reinforcement ratio is computed as the total area of steel reinforcement divided

123

SW

SW

Carapelle

Vallone Volpi

DB + SH (tr),
DB + SR
(long)

SH (transv),
SH + SR
(long)
DB + SH (tr),
DB + SR
(long)
NP, SB

NP, SB

33.5

33.5

33.5

33.5

33.5

33.5

33.5

33.5

33.5

Span length
(m)

44

11

25

26

36

Variability of
heights [VH]
(%)

0.44

0.30

1.38

1.38

0.55

0.55

0.24

0.32

0.44

Long. reinforc.
ratio [l ] (%)

a SF simple frame, SS single shaft, SW single wall


b SH steel hinges, DB dowel bars, SP steel pendulum, SR steel rollers, SB sliding bearings, NP neoprene pads

SF

Serra Lupi

NP

SS

SF

Lauretta

SW

SS

Castello

Ceraso

Macchione

NP

SW

San Gennaro

SH (transv),
SH + SP
(long)
NP, SB

SF

Fiumarella

Bearing typesb

Pier typesa

Bridges

Table 3 Main characteristics of the analyzed bridges

0.15

0.085

0.11

0.11

0.16

0.16

0.29

0.07

0.14

Trasv.
reinforc. ratio
[s ] (%)

3540

3540

3540

3540

3540

3540

4045

3540

3540

Concrete
strength
( MPa)

440

270

230

230

440

440

270

230

270

Steel yielding
( MPa)

Bull Earthquake Eng (2011) 9:14631498


1479

123

1480

Bull Earthquake Eng (2011) 9:14631498


DS11
DS21
DS31

Transverse

Lauretta

Sa (g)

0.5
0.4

1
0.75

0.3

0.5

0.2
0.1

Sd (mm)

0.25

DS35

PGA (g)

50

100

150

200

Sa (g)

0.4
0.3

Ceraso

0.75
0.5

0.2
0.1

Sd (mm)

0.2

0.4

0.6

0.8

P(DPL|PGA)

0.5

0.25

PGA (g)

50

100

200

San Gennaro

0.75
0.5

0.2
Sd (mm)
0

0.2

0.4

0.6

0.8

P(DPL|PGA)

0.4

150

Sa (g)

0.6

0.25

PGA (g)

0
0

50

100

150

200

250

300

0.6

Sa (g)

0.75
0.5

Sd (mm)
0

0.2

0.4

0.6

0.8

P(DPL|PGA)

Castello

0.2

0.25

PGA (g)

0
100

200

0.6

300

Carapelle

Sa (g)

0.75
0.5

0.2
Sd (mm)
0

0.2

0.4

0.6

0.8

P(DPL|PGA)

0.4

DS34

0
0

0.4

DS13
DS23
DS33

P(DPL|PGA)

Longitudinal

DS12
DS22
DS32

0.25

PGA (g)

0
0

100

200

300

0.2

0.4

0.6

0.8

Fig. 10 (Left) capacity spectra and performance points of the bridges in the longitudinal and transverse directions. (Right) fragility curves associated to the main damage states registered during the pushover analyses

by the gross area of the pier cross section. The transverse reinforcement ratio is computed as
the volume of transverse reinforcement divided by the volume of concrete between a single
spacing. Different types of bearing devices have been found, which basically realise three
different types of pier-deck connections, i.e.: (i) fixed (steel hinges), (ii) moveable (steel

123

Bull Earthquake Eng (2011) 9:14631498

1481
DS11
DS21
DS31

Transverse
Longitudinal

0.75

0.6

0.5

0.3

0.25

Sd (mm)

DS34

DS35

P(DPL|PGA)

Fiumarella

DS13
DS23
DS33

PGA (g)

Sa (g)

1.2

50

0.9

100

Serra Dei Lupi

0.75

0.6

0.5

0.3

0.25

Sd (mm)

0.2

0.4

0.6

0.8

P(DPL|PGA)

PGA (g)

Sa (g)

1.5

50

100

0.2

Vallone Delle Volpi

0.75
0.5

0.5
Sd (mm)
0

0.4

0.6

0.8

P(DPL|PGA)

0.9

Sa (g)

1.2

DS12
DS22
DS32

0.25

PGA (g)

50

100

Macchione

Sa (g)

1.2
0.9

0.75

0.6

0.5

0.3

0.25

Sd (mm)

0.2

0.4

0.6

0.8

P(DPL|PGA)

PGA (g)

0
0

50

100

0.2

0.4

0.6

0.8

Fig. 10 Continued

pendulum, steel rollers and sliding bearings) and (iii) semirigid (neoprene pads and dowel
bars) connections.
All the selected bridges are multi-span simply supported deck bridges, with span lengths
of approximately 33 m.
All the bridges taken into consideration were built between 1969 and 1971, according to
the pre-1971 Italian Regulations for RC structures. As a consequence, they were not designed
to resist seismic loads, though being built in moderate-to-high seismicity regions. In addition, neither specific design criteria (capacity design) nor specific rules for seismic detailing
(minimum amount of reinforcement, maximum stirrup spacing at the beam and column ends,
etc.) were followed. The seismic vulnerability of the selected bridges is then expected to be
high.

123

1482

Bull Earthquake Eng (2011) 9:14631498

3.2 Application of the proposed procedure


The bridge structures selected have been modelled according to the modelling assumptions
discussed before (see Sect. 2.2) using the software package SeismoStruct (SeismoSoft 2006),
in which a routine for the DAP analysis of structures is implemented. The DAP analysis has
been carried out, separately in the transverse and longitudinal direction of the bridge. In
Fig. 10 the capacity curves relevant to each bridge structure are shown and the Damage
States (DSij ) considered in the application of the IACSM are identified.
In Table 4 the Damage States (DSij ), experienced by each bridge structure in the longitudinal and transverse direction, are described and the corresponding PGA values, derived
with IACSM using the elastic spectrum of Eurocode 8 for soil type B (CEN 2005) as demand
spectrum, are reported. As can be seen, in the longitudinal direction (see Table 4) the PGA
values corresponding to the attainment of the first slight damage state (PL1) range between
approximately 0.05 and 0.25 g and those corresponding to the first severe damage state (PL3)
between approximately 0.2 and 1 g. In the transverse direction (see Table 4), instead, the
PGA values corresponding to the attainment of the first slight damage state (PL1) range
between approximately 0.16 and 0.7 g and those corresponding to the first severe damage
state (PL3) between approximately 0.4 and 1 g. The examination of the damage states in
the longitudinal direction (see Table 4) clearly points out that the behaviour of joints and
abutments can considerably affect the seismic response of multi-span simply supported deck
bridges. Joints and abutments must be correctly modelled, in order to properly estimate the
seismic vulnerability of existing bridges, especially when the seismic response of the bridge
is governed by the displacement capacity of the bearing devices.
In the last column of Table 4, the PGA values on stiff soil, at the bridge site, having 10% probability of exceedance in 50 years are reported. The ratio between PGAPLi
and PGA10%/50y provides a first (deterministic) measure of the seismic vulnerability of
the bridge. As can be noted, in most of the cases, the values of PGA10%/50y result lower
than PGAPL3 . The only exceptions are represented by Macchione, Fiumarella and Serra
dei Lupi bridges in the longitudinal direction, where the values of PGA10%/50y result very
close to PGAPL3 , indicating that extensive structural damage is expected to occur in case
of ground motions with 500 years mean return period. Actually, a number of seismic retrofit measures, consisting in the strengthening of the piers and in the replacement of the
bearing devices, are being undertaken for the three bridges under consideration. It is worth
to observe that for all the bridges examined, with the exception of Vallone delle Volpi and
Candelaro in the transverse direction, the values of PGA10%/50y result greater than PGAPL1 ,
these latter being associated to either pier yielding or joint closure or failure of neoprene
pads.
On the left-hand side of Fig. 10, the adaptive capacity spectra of the bridges under consideration are shown, separately in the longitudinal and transverse direction. On each capacity
curve, the PPs corresponding to the DSs registered during pushover analysis are identified.
The fragility curves associated to the most important DSs (one for each PL) are reported on
the right-hand side of Fig. 10.
As a matter of fact, three classes of structures can be identified: (i) bridges with single
shaft piers and neoprene pads (e.g. Lauretta and Ceraso bridges), whose seismic response is
governed by pier yielding, (ii) bridges with wall piers and sliding bearings combined with
neoprene pads (e.g. Carapelle, Castello and San Gennaro bridges), whose seismic response
is governed by the mechanical behaviour of the bearing devices and the clearance of the
joints, (iii) bridges with frame piers and steel hinges combined with sliding/roller bearings
(e.g. Fiumarella, Macchione, Serra dei Lupi and Vallone delle Volpi bridges), whose seismic

123

Bull Earthquake Eng (2011) 9:14631498

1483

Table 4 Performance levels, damage states and corresponding PGA values in the longitudinal and transverse
direction of the examined bridges
Bridge

PL

Longitudinal

Transverse

PGA
10%/50y

DS
Lauretta

Ceraso

PL1 DS11: closure of


abt-deck joint
DS12: pier yielding
PL2 DS21: 50% ultimate
ductility of pier
PL3 DS34: passive
resistance of abutment
PL1 DS11: closure of
abt-deck joint
DS12: pier yielding

PL2 DS21: 50% ultimate


ductility of pier
PL3 DS32: pier ultimate
ductility
San Gennaro PL1 DS11: closure of
abt-deck joint
DS13: failure of
neoprene pads
PL2

Castello

PGA (g) DS

PGA (g)

0.18

DS12: pier yielding

0.25

DS21: 50% ultimate


ductility of pier
DS32: pier ultimate
ductility
DS12: pier yielding

0.38

0.28
0.50
0.60
0.25

Fiumarella

0.45
0.54
0.093

0.23

0.34

DS21: 50% ultimate


ductility of pier
DS32: pier ultimate
ductility
DS11: closure of
deck-deck joint

0.34
0.42
0.18

0.25

0.30

PL3 DS34: passive


0.32
resistance of abutment
PL1 DS11: closure of
0.073
abt-deck joint

PL3 DS34: passive


0.37
resistance of abutment
PL1 DS11: closure of
0.078
deck-deck joint
DS13: failure of
0.25
neoprene pads
PL2 PL3 DS34: passive
0.48
resistance of abutment
PL1 DS12: pier yielding
0.15
DS11: closure of
deck-deck joint
PL2 DS21: 50% ultimate
ductility of pier
DS23: active resistance
of abutment
DS22: Displ. capacity
of steel pendulum
PL3 DS32: pier ultimate
ductility

0.50

0.35

PL2 -

Carapelle

0.34

DS22: Displ. capacity


of sliders
DS35: deck unseating

0.48

DS11: closure of
deck-deck joint
DS13: failure of
neoprene pads
DS22: Displ. capacity
of sliders
DS35: deck unseating

0.18

DS11: closure of
deck-deck joint
DS13: failure of
neoprene pads
DS22: Displ. capacity
of sliders
DS35: deck unseating

0.23

0.70
0.25

0.40
0.49
0.82

0.39

0.25

0.40
0.68

DS12: pier yielding

0.22

DS21: 50% ultimate


ductility of pier

0.45

DS32: pier ultimate


ductility

0.60

0.33

0.16
0.25
0.26
0.28
0.33

123

1484

Bull Earthquake Eng (2011) 9:14631498

Table 4 Continued
Bridge

PL

Longitudinal

Transverse

PGA
10%/50y

DS

PGA (g) DS

Serra dei Lupi PL1 DS11: closure of


abt-deck joint
PL3 DS33: pier shear failure

0.19

0.30

DS33: pier shear failure

Vallone Volpi

PL1 DS12: pier yielding

0.13

DS11: closure of
deck-deck joint
PL2 DS23: active resistance
of abutment
PL3 DS31: failure of dowel
bars
PL1 DS11: closure of
deck-deck joint
DS12: pier yielding

0.18

0.36

0.58

DS31: failure of dowel


bars
DS11: closure of
deck-deck joint
DS12: pier yielding

Macchione

PL2 DS22: Displ. capacity


of steel roller
DS21: 50% ultimate
ductility of pier
PL3 DS35: deck unseating

0.05
0.13
0.10

PGA (g)
0.33
0.37
0.26

1.04
0.16

0.33

0.30

DS21: 50% ultimate


ductility of pier

0.44

DS32: pier ultimate


ductility

0.60

0.29
0.31

response is strongly influenced by the arrangement of the pier-deck connections along the
bridge.
As expected, for the first class of bridges the distance between the fragility curves associated to PL1 and PL3 increases according to the available ductility (see Fig. 10). On the
contrary, for the second class of bridges the distance between the fragility curves associated to PL1 and PL3 strongly depends on the displacement capacity of the bearing devices
and, in the longitudinal direction, it is strongly conditioned by the closure of the joints with
consequent attainment of the passive resistance of the abutment-backfill system (see Fig. 10).
Finally, for the third class of bridges, the distance between the fragility curves associated
to PL1 and PL3 mostly depends on the flexural/shear strength of the piers with steel hinge
pier-deck connections.
Comparing the fragility curves relevant to bridges of the same class (e.g. Lauretta and
Ceraso, Carapelle and Castello, Fiumarella and Serra dei Lupi), it turns out that, despite an
apparent similarity reflecting the corresponding similarities in the seismic response of the
bridge, the development of the damage states, as well as the PGA values associated to each
PL, can differ significantly among bridges of the same class, due to differences in either geometric characteristics (pier heights, joint clearance, etc.) or mechanical characteristics (pier
reinforcement ratios, bearing device strength, etc). This emphasizes the importance of the
proposed methodology, revealing that it is tricky to define fragility curves representative of
entire classes of bridges based on typological aspects only, as done in some previous studies
(Basz and Mander 1999; Nielson and DesRoches 2007).
Another interesting aspect is the comparison between the two bridge directions, in the
attempt to identify a weaker direction, which, if actually existed, would significantly reduce
the computational effort for deriving fragility curves, since only one set of analyses would
be required (Moschonas et al. 2009). The results of this study show that the critical direction

123

Bull Earthquake Eng (2011) 9:14631498

1485

mainly depends on the class of bridges considered. For the bridges of the first class, the
weaker direction is the transverse direction, if the definition of the weaker direction is based
on the higher two PLs (i.e. PL2 and PL3, respectively), with PGA values by 2030% lower
than in the longitudinal direction. This can be ascribed to the higher equivalent damping
ratio of the bridge in the longitudinal direction, due to the nonlinear hysteretic behaviour
of all the piers, while in the transverse direction yielding is limited to one or two piers
only.
For the bridges of the second class, instead, the weaker direction is the longitudinal direction, with PGA values practically half those in the transverse direction. The main reason
for the longitudinal direction being the critical one is that the gaps between adjacent decks
close relatively early, leading to the failure of the bridge due to the attainment of the passive
resistance of the abutment-backfill system before pier yielding.
For the bridges of the third class, finally, the weaker direction is once again the longitudinal direction, with PGA values significantly lower than those in the transverse direction, for
all the PLs considered. The main reason for that is the presence of steel hinges and dowel
bar connections in the longitudinal direction, where the bridge results more vulnerable, due
to the lower out-of-plane strength of wall and frame piers.
In Table 5 the values of the risk index derived from the fragility curves shown in
Fig. 10, using the newest Italian earthquake hazard map, developed by the National Institute of Geophysics and Vulcanology (INGV), are reported. As can be seen, in the longitudinal direction, the risk index ranges from 2.8 102 to 9.2 102 for PL1, and
from 0.4 102 to 2.5 102 for PL3. Considering PL3, the bridges with the highest risk index are Serra Dei Lupi, Macchione and Fiumarella bridges, characterised by
frame piers and steel hinges combined with pendulum bearings. In the transverse direction, the risk index ranges from 2 102 to 3.6 102 for PL1, and from 7 104 to
1.9 102 for PL3. Considering PL3, the bridges with the highest risk index are Ceraso
(single shaft piers and neoprene pads) and Serra dei Lupi (frame piers with steel hinges and
rollers).
3.3 Comparison with nonlinear response time-history analysis results
Comprehensive Nonlinear response Time-History Analyses (NTHA) have been carried out
to evaluate the accuracy of the IACSM, employed in the proposed methodology to estimate the PGA values associated to given DSs. The bridge models for NTHA have been
implemented in SAP2000_Nonlinear (Computers and Structures 1999), adopting the same
modelling assumptions made within the proposed procedure (see Sect. 2.2). The NTHA have
been performed using a set of 7 artificial accelerograms strictly compatible, on average, with
the 5%-damped acceleration response spectrum provided by Eurocode 8 for soil type B (CEN
2005) (see Fig. 11). The input ground motions have been scaled (on average) to the PGA
values provided by IACSM for a number of characteristic DSs (see Table 5). A total of 280
NTHA (9 bridges 2 directions 7 accelerograms 13 damage states) have been carried out.
The accuracy of the proposed procedure has been evaluated by comparing (for each DS) the
bridge displacement profile expected based on IACSM with the envelope of the maximum
bridge displacements (averaged on 7 accelerograms) obtained from NTHA.
In Fig. 12, IACSM predictions and NTHA results are compared for the longitudinal
direction of the Ceraso bridge, characterised by single shaft piers and neoprene pier-deck
connections (see Table 3). The comparison is made in terms of absolute displacements of the
decks and relative displacements of the joints relevant to three different DSs, corresponding

123

1486

Bull Earthquake Eng (2011) 9:14631498

Table 5 Seismic risk index of the selected bridges for each performance level and seismic direction
Bridge (Pier type/bearing type)

Lauretta (SS/NP)

Ceraso (SS/NP)

San Gennaro (SW/NP + SB)

Castello (SW/NP + SB)

Carapelle (SW/NP + SB)

Fiumarella (SF/SH + SP, SH)

Serra Dei Lupi (SF/DB + SR, DB + SH)


Vallone Delle Volpi (SW/DB + SR, DB + SH)

Macchione (SF/SH + SR, SH)

Longitudinal

Transverse

PLs

RL (102 )

PL1

4.752

PL1

3.288

PL2

1.074

PL2

1.779

PL3

0.729

PL3

1.074

PL1

3.288

PL1

3.642
2.131

PLs

RT (102 )

PL2

1.317

PL2

PL3

0.917

PL3

1.495

PL1

4.978

PL1

2.728

PL2

0.543

PL3

1.196

PL3

0.213

PL1

5.554

PL1

2.516

PL2

0.393
0.088

PL3

0.749

PL3

PL1

5.526

PL1

1.993

PL2

0.794

PL3

0.543

PL3

0.231

PL1

5.534

PL1

3.793

PL2

3.255

PL2

1.318

PL3

2.218

PL3

0.735

PL1

4.445

PL3

2.551

PL3

1.853

PL1

4.157

PL2

1.089

PL3

0.399

PL3

0.073

PL1

9.207

PL1

5.235

PL2

7.302

PL2

1.374

PL3

2.434

PL3

0.735

to pier yielding (PGA = 0.35 g), attainment of 50% pier ultimate ductility (PGA = 0.45 g)
and pier collapse (PGA = 0.55 g), respectively.
As can be seen, two series of results are reported for IACSM. They differ in the way of
application of the lateral displacements in the DAP analysis: from left to right (IACSM )
and from right to left (IACSM ), respectively. The execution of two pushover analyses is
fundamental for bridges with unsymmetrical configuration of piers and/or pier-deck connections. In these situations, reference to the envelope between IACSM and IACSM should
be made. For the sake of brevity, in the following the acronym IACSM is used to identify the
envelope between IACSM and IACSM .
As far as joint displacements are concerned, it should be noted that negative values correspond to two adjacent decks that move closer together. The limit value in this case is
represented by the joint gap and corresponds to pounding between the decks.
Generally speaking, a good agreement between IACSM predictions and NTHA results is
observed for the Ceraso brigde, for all the DSs considered. Indeed, percent errors, calculated

123

Bull Earthquake Eng (2011) 9:14631498

1487

Spectral Acceleration (m/s 2 )

Time Series
EC8 Spectrum

0
0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

Spectral Displacement (m)


Fig. 11 Comparison between the EC8 soil B response spectrum ( = 5%) and the response spectra derived
from the input accelerograms

as the absolute difference between NTHA and IACSM results divided by the exact NTHA
value, less than 14% are found in terms of deck displacements, with the IACSM tending
to be a little conservative. Moreover, the level of accuracy of the IACSM is practically the
same passing from slight damage states (DS12 in Fig. 12), where all the piers are still elastic, to severe damage states (DS32 in Fig. 12), where extensive plastic deformations occur.
The good accordance between the two displacement profiles related to DS32 confirms that
the proposed procedure is able to capture the actual distribution of plastic deformations in
the structure. In particular, both IACSM and NTHA identify pier n. 1 (in the longitudinal
direction) and pier n. 4 (in the transverse direction) as the critical members of the bridge,
where yielding, 50% ultimate ductility and collapse first take place. Also piers n. 2 and n. 3,
however, undergo extensive plastic deformations in both directions.
As far as joint displacements is concerned, significant differences between IACSM predictions and NTHA results are observed. They can be ascribed to asynchronous movements
between adjacent decks due to the higher-modes response of the bridge, that appears to be a
little underestimated by the DAP analysis.
Similar observations can be made examining the seismic response in the longitudinal direction of the Carapelle (see Fig. 13) and Vallone delle Volpi bridges (see Fig. 14), differing
from the Ceraso bridge in pier type and pier-deck connections (see Table 3). In the first case,
the collapse of the bridge (PL3) is caused by the attainment of the passive resistance of the
abutment at 0.48 g PGA, after failure of neoprene pads (PL1) at 0.25 g PGA. No damage is
registered for piers. In the second case, the damage first occurs in the piers (PGA = 0.13 g)
and subsequently evolves in abutment (attainment of the active resistance at 0.36 g PGA)
and finally in the bearings, with the failure of some dowel bar connections at 0.58 g PGA.
Referring to DS31, plastic deformation occurs for both piers of the Vallone delle Volpi bridge,
with ductility demand of 25% of the ultimate ductility.
In both the cases, a good accordance between IACSM predictions and NTHA results is
observed in terms of deck displacements, with errors ranging from 1 to 20% and from 7
to 30%, for the Carapelle and Vallone delle Volpi bridge, respectively. The importance of
the dual analysis (i.e. IACSM and IACSM ) is apparent for the Carapelle bridge, whose
seismic response under strong earthquakes is dominated by the abutment behaviour, which
determines a not symmetrical response of the bridge. The presence of sliding/roller bearings

123

1488

Bull Earthquake Eng (2011) 9:14631498


CERSASO BRIDGE A16 NAPOLI-CANOSA (km 73+385)
D1

D2

J1

D3

J2

D4

J3

D5

J4

A1

NP

A2
P1

200
150

P4

SS

150

D deck (mm)

250

P3

P2

NTHA
IACSM
IACSM

100
50

-50

50

-100

-150

250
200
150

D2

D3

D4

D deck (mm)

D1

d joint (mm)

0
100

D5
NTHA
IACSM
IACSM

DS12 (0.35 g)
J1

J2

J3

J4

150
100
50

-50

50

-100

150

D2

D3

D4

D deck (mm)

200

DS21 (0.45 g)

-150
D1

250

d joint (mm)

0
100

D5
NTHA
IACSM
IACSM

J1

J2

J3

J4

150
100
50

-50

50

-100

d joint (mm)

0
100

DS32 (0.54 g)

-150
D1

D2

D3

D4

D5

J1

J2

J3

J4

Fig. 12 Seismic response of the Ceraso bridge in the longitudinal direction. Comparison between IACSM
predictions and NTHA results (maximum values averaged over 7 ground motions): (left) absolute deck displacements and (right) relative joint displacements

in the longitudinal direction of the Carapelle and Vallone delle Volpi bridge considerably
emphasises the effects of the higher modes, leading to highly irregular profile envelopes
of dynamic joint displacements, for all the levels of seismic excitation considered, which
are significantly underestimated by the IACSM. Nevertheless, the IACSM is consistently
capable of reproducing the correct trend of the NTHA envelopes.

123

Bull Earthquake Eng (2011) 9:14631498

1489

CARAPELLE BRIDGE A16 NAPOLI-CANOSA (km 11+55)


A1

D1
SB

D2

J1
NP

D3

J2
NP

SB

SB

D4

J3
NP

SB

SB

SW

P3

P2

100
NTHA
IACSM
IACSM

DS34 (0.48 g)
50

100

50

-50

d joint (mm)

Ddeck (mm)

150

NP

P4

P1

200

NP/SB

A2

D5

J4
NP

-100
D1

D2

D3

D4

D5

J1

J2

J3

J4

Fig. 13 Seismic response of the Carapelle bridge in the longitudinal direction. Comparison between IACSM
predictions and NTHA results (maximum value averaged over 7 ground motions): (left) absolute deck displacements and (right) relative joint displacements

In Fig. 15, IACSM predictions and NTHA results are compared for the transverse direction of the Ceraso bridge. The comparison is made in terms of deformed shapes of the decks
associated to three different DSs, corresponding to pier yielding (PGA = 0.23 g), attainment
of 50% pier ultimate ductility (PGA = 0.34 g) and pier collapse (PGA = 0.42 g), respectively.
Based on the IACSM, the critical pier, where yielding, 50% ultimate ductility and collapse
first take place, is the pier n. 3 (see Fig. 15). Also pier n. 2 undergoes extensive plastic
deformations while piers n. 1 and pier n. 4 remain elastic.
For the sake of clarity, in Fig. 15 transverse displacements and bridge coordinates are
reported in two different scales. From a graphical point of view, this determines a distortion
of the deformed shape of the bridge that considerably amplifies the rotations of the decks.
Nevertheless, the comparison between expected and actual deformed shapes clearly points
out the great accuracy of the IACSM in the prediction of the PGA values associated to slightto-severe damage states. Indeed, the percent errors in the evaluation of the actual maximum
deck displacements do not exceed 14% for DS12 and 20% for DS32 and, on average, they
result of the order of 9% for DS12 and of the order of 13% for DS32. The NTHA results also
confirm that pier n. 3 is the critical element of the bridge with ductility demands that differ
from those expected based on IACSM less than 8%, regardless the DS considered.
Similar considerations apply also for different bridge configurations, characterized by
either highly irregular layout of pier heights (like the Macchione bridge examined in Fig. 16)
or different pier-deck connections at the ends of the same span (like the San Gennaro bridge
of Fig. 17), for which significant higher mode contributions are expected. The structural
behavior of the Macchione bridge (Fig. 16) changes substantially passing from DS12 (yielding of piers n. 5 and 8) to DS32 (collapse of pier n. 6), due to a redistribution of seismic
forces between the piers. The seismic response of the San Gennaro bridge (Fig. 17), on the
other hand, is governed by the displacement capacity of the sliding bearings, with wall piers
that remain elastic. As a result, the three decks of the San Gennaro bridge tend to move

123

1490

Bull Earthquake Eng (2011) 9:14631498

VALLONE DELLE VOLPI BRIDGE A16 NAPOLI-CANOSA (km 121+786)

A1

D1
DB

D2

J1
SR

DB

SR

NTHA
IACSM
IACSM

100

DB

SW

50

-100
0

-150
D1

D2

D3

J1

Ddeck (mm)

100

DS23 (0.36 g)

50

-50
-100
0

-150
D1

D2

d joint (mm)

50

D3

J1

J2

150

Ddeck (mm)

NTHA
IACSM
IACSM

100

100

DS31 (0.58 g)

50
0
-50

50
-100
0

-150
D1

D2

D3

d joint (mm)

150

J2

150
NTHA
IACSM
IACSM

100

200

d joint (mm)

0
-50

150

DS12 (0.13 g)

100

50

200

DB/SR

150

Ddeck (mm)

150

DB
P2

P1
200

A2

D3

J2

J1

J2

Fig. 14 Seismic response of the Vallone Delle Volpi bridge in the longitudinal direction. Comparison between
IACSM predictions and NTHA results (maximum values averaged over 7 ground motions): (left) absolute deck
displacements and (right) relative joint displacements

independently of each other. Nevertheless, in both the cases, a good accordance between the
two deformed shapes is observed, with errors that, on average, do not exceed 25% for the
Macchione bridge and 10% for the San Gennaro bridge.

123

Bull Earthquake Eng (2011) 9:14631498

1491

CERASO BRIDGE A16 NAPOLI-CANOSA (km 73+385)


D1

D2

J1

D3

J2

D4

J3

D5

J4

A1

A2

200

Ddeck (mm)

P1

P2

P3

P4

NP

SS

DS12 (0.23 g)
NTHA
IACSM

100

Ddeck (mm)

200

D1

D2

D3

D4

D5

DS21 (0.34 g)
NTHA
IACSM

100

Ddeck (mm)

200

D1

D2

D3

D4

D5

DS32 (0.42 g)
NTHA
IACSM

100

D1

D2

D3

D4

D5

Fig. 15 Seismic response of the Ceraso bridge in the transverse direction. Comparison between IACSM
predictions and NTHA results (maximum values averaged over 7 ground motions) in terms of deformed shape
of the bridge

To better measure the accuracy of the IACSM in capturing the exact maximum deformed
shape of the bridge and the exact maximum pier displacement profile, two indices have been
computed for each DS selected. The first index is the so-called Bridge Index (BI), already
used by Pinho et al. (2007) for similar comparisons. It is defined as:


Dj,IACSM
BI = medianj=1,...,2Nd
(20)
Dj,NTHA
where Dj,IACSM is the displacement of the jth deck end provided by IACSM, Dj,NTHA is the
corresponding maximum displacement (averaged between 7 accelerograms) derived from
NTHA and Nd is the number of decks.

123

1492

Bull Earthquake Eng (2011) 9:14631498


MACCHIONE BRIDGE A16 NAPOLI-CANOSA (km 91+555)
SH/SR

A1

D1

J1

D2
P1

D3

J2

P2

J3

D4
P3

J4

D5
P4

J5

D6

J6

D7

D8

J7

D9

J8

J9

D10
A2

P6

P5

P7
P8

P9

SF

100

Ddeck (mm)

150
NTHA
IACSM

DS12 (0.30 g)

50

150

100

Ddeck (mm)

D1

D2

D3

D4

D5

D6

D7

D8

D9

D1 0
NTHA
IACSM

DS21 (0.44 g)

50

100

Ddeck (mm)

150

D1

D2

D3

D4

D5

D6

D7

D8

D9

D1 0

NTHA
IACSM

DS32 (0.60 g)

50

D1

D2

D3

D4

D5

D6

D7

D8

D9

D1 0

Fig. 16 Seismic response of the Macchione bridge in the transverse direction. Comparison between IACSM
predictions and NTHA results (maximum values averaged over 7 ground motions) in terms of deformed shape
of the bridge

The second index is the Pier Index (PI), defined as:



PI = medianj=1...2Np

di,IACSM
di,NTHA


(21)

where di,IACSM is the displacement of the ith pier provided by IACSM, di,NTHA is the corresponding maximum displacement (averaged between 7 accelerograms) derived from NTHA
and Np is the number of piers.
In addition, a third index, referred to as Critical Element Index (CEI), has been calculated
to evaluate the ability of IACSM in capturing the actual seismic demand to the critical
element (pier, bearing device, abutment and joint) of the bridge predicted by the IACSM.

123

Bull Earthquake Eng (2011) 9:14631498

1493

SAN GENNARO BRIDGE A16 NAPOLI-CANOSA (km 126+371)

A1

D1
SB

D2

J1
NP

SB

NP

A2

D3

J2
NP

NP/SB

SB

SW

200
NTHA

DS22 (0.48 g)

150

50
0

Ddeck (mm)

100

IACSM

D1

D2

D3

Fig. 17 Seismic response of the San Gennaro bridge in the transverse direction. Comparison between IACSM
predictions and NTHA results (maximum values averaged over 7 ground motions) in terms of deformed shape
of the bridge

The CEI index is defined as:


CEI =

cr,IACSM
cr,NTHA

(22)

where cr,IACSM is the displacement of the critical element of the bridge predicted by the
IACSM and cr,NTHA the corresponding maximum displacement (averaged between 7 accelerograms) derived from NTHA.
In Table 6 the values of BI, PI and CEI for the selected DSs are reported, separately for
the transverse and longitudinal direction of each bridge structure. The ideal target value of
BI, PI and CEI is always 1.
In the longitudinal direction, the BI ranges between 0.87 and 1.18, the PI index between
0.75 and 1.22 and the CEI between 0.87 and 1.20. In 80% of the cases examined, both BI
and CEI result between 0.9 and 1.1. The percentage reduces to 65% for the PI.
In the transverse direction, excluding a few cases, the BI ranges between 0.89 and 1.05,
the PI index between 0.48 and 1.05 and the CEI index between 0.90 and 1.31. In 75% of the
cases considered, the BI, PI and CEI result between 0.9 and 1. The low values of PI for the
DS22 of the bridges with shear walls and sliding bearings in the longitudinal direction (i.e.
San Gennaro, Castello and Carapelle bridges) must be ascribed to the order of magnitude
of the top pier displacements (less than 1 mm), the response of this class of bridges being
governed by the large displacements of sliding bearings with piers that remain elastic.
The lowest value of BI in the transverse direction is related to the DS32 of the Fiumarella
bridge, for which the IACSM significantly underestimates the actual maximum deformed
shape of the bridge, since it does not predict correctly the inelastic behaviour of pier n. 1 and
n. 3 (see Fig. 18). Nevertheless, the ductility demands to the critical pier are captured with
good accuracy (see Fig. 18).
The highest value of BI in the transverse direction is registered for the DS22 of the
Castello bridge. In this case the IACSM significantly overestimates the actual maximum
deformed shape of the bridge, which is mainly governed by the friction behaviour of the
sliding bearings (see Fig. 19). The main reason for this inconsistency could be the underes-

123

1494

Bull Earthquake Eng (2011) 9:14631498

Table 6 Bridge Index (BI), Pier Index (PI) and Critical Element Index (CEI) of the bridges under examination
for different damage states
Bridge

Longitudinal
DSs

Lauretta (SS/NP)

Ceraso (SS/NP)

San Gennaro (SW/NP + SB)

Castello (SW/NP + SB)

Carapelle (SW/NP + SB)

Fiumarella (SF/SH + SP, SH)

BI

Transverse

PI

CEI

DSs

BI

PI

CEI

DS12

1.03

0.99

0.99

DS12

1.06

1.05

1.07

DS21

0.92

0.85

1.15

DS21

0.98

1.07

DS34

0.91

0.83

0.91

DS32

0.91

0.89

0.91

DS12

0.95

0.95

0.93

DS12

0.9

0.95

0.92

DS21

1.07

0.99

1.10

DS21

0.95

0.89

0.98

DS32

1.04

0.96

1.10

DS32

0.88

0.92

0.96

DS11

0.87

0.82

0.87

DS11

1.01

1.02

1.01

DS22

1.06

0.48

1.06

DS34

1.02

1.02

1.05

DS11

0.98

0.86

0.98

DS11

1.02

1.02

1.02

DS22

1.45

0.77

1.31

DS34

1.11

0.8

1.11

DS11

0.95

0.97

0.93

DS11

1.03

0.99

1.03

DS22

0.86

0.65

0.74

DS34

1.06

0.74

1.02

DS12

0.94

0.94

0.94

DS12

0.95

1.04

DS21

1.10

1.03

1.03

DS21

0.87

0.78

1.17
0.92

DS32

1.10

1.03

1.03

DS32

0.61

0.31

Serra Dei Lupi (SF/DB + SR, DB + SH)

DS33

1.10

1.11

1.2

DS33

1.01

1.01

1.01

Vallone Delle Volpi (SW/DB + SR,


DB + SH)

DS12

1.18

1.08

0.98

DS23

0.94

1.10

0.94

DS31

1.14

1.22

1.11

DS31

0.90

0.90

0.90
0.96

Macchione (SF/SH + SR, SH)

DS12

0.99

0.99

DS21

0.98

0.98

1.24

DS35

1.06

1.07

0.99

DS32

0.93

0.93

1.20

timation of the energy dissipation capacity of the bridge ( P P 40%), as a consequence


of the damping reduction factor employed [see Eq. (15)]. Significant improvements in the
accuracy of the results can be reached simply by removing the limitation > 0.5 (see
Fig. 19).

4 Conclusions
A new methodology for the seismic evaluation of multi-span simply supported bridges has
been proposed. The proposed methodology is based on an Inverse Adaptive application of
the Capacity Spectrum Method. For this reason the acronym IACSM is used to identify
the proposed methodology. The IACSM provides the PGA values associated to predefined
Performance Points of the structure corresponding to different Damage States of the critical members of the bridge (piers, abutments, bearing devices, joints). Based on these PGA

123

Bull Earthquake Eng (2011) 9:14631498

1495

FIUMARELLA BRIDGE A16 NAPOLI-CANOSA (km 95+0.93)


D1

D2

J1

D3

J2

D4

J3

P2

d p (mm)

0
-25

P3

4000

4000

-50

8000

Fp (kN)

Fp (kN)

Fp (kN)

8000

SF

P3

P2

P1

P1

SH
A2

A1

25

-4000

50

4000
d p (mm)

0
-50

-25

25

-4000

NTHA

8000

50

-50

-25

-4000

NTHA

IACSM

25

50
NTHA
IACSM

IACSM

-8000

d p (mm)

-8000

-8000

Fig. 18 Seismic response of the Fiumarella bridge in the transverse direction at 0.60g PGA (DS32). Comparison between IACSM predictions and NTHA results (accelerogram n. 3) in terms of cyclic behaviour of
piers
CASTELLO BRIDGE A16 NAPOLI-CANOSA (km 582+931)

A1

D1
SB

A2

D2

J1
NP

NP
P1

NP/SB

SB

SW

200
PGA=0.65g

100
50

Ddeck (mm)

150

IACSM
NTHA =()
NTHA =() 0.55

PGA=0.49g

D1

D2

Fig. 19 Seismic response of the Castello bridge in the transverse direction (DS22). Effects of the damping
reduction factor on the accuracy of the results

values, a number of fragility curves are derived to describe the seismic vulnerability of the
bridge from a probabilistic point of view. The seismic risk is then computed as convolution
integral of the product between the seismic vulnerability of the bridge (expressed by the
fragility curves associated to selected Damage States) and the seismic hazard of the bridge
site (expressed by a suitable seismic hazard curve).
The IACSM has been applied to a set of nine multi-span simply supported deck bridges
of the Italian A16 NapoliCanosa highway, differing in pier types, pier layout and pier-deck
connections. The predictions of the IACSM have been compared with the results of Nonlinear
response Time-History Analyses (NTHA), carried out using a set of seven accelerograms,

123

1496

Bull Earthquake Eng (2011) 9:14631498

compatible with the EC8-soilB response spectrum scaled to the PGA values provided by
IACSM for selected Damage States. The comparison has been made in terms of maximum
deformed shapes of the deck, joint displacements and top pier displacements.
The comparison between IACSM predictions and NTHA results confirms the good accuracy of the proposed methodology in predicting the PGA values associated to slight-to-severe
Damage States. In all the examples of application considered, indeed, the IACSM correctly
identified the critical element of the bridge, where first a given damage state was reached.
The displacement profile of the bridge predicted by the IACSM (including joint displacements, top pier displacements, bearing device displacements and deck rotations) is in good
agreement with the maximum deformed shape of the bridge found with NTHA.
Although the proposed methodology appears very promising, there are a number of aspects
that require further investigation. In particular, the number of bridges considered was too small
to examine the effectiveness of the method to address directional effects and soil-structure
interaction effects. Curved bridges and skewed bridges, moreover, have not yet been studied.
A related area of research will involve an investigation as to when the assumption of a rigid
superstructure is valid, as such an assumption greatly simplifies the process for evaluating
the inelastic displacement pattern of the bridge. Future research should focus also on the
influence of different modeling approaches, more refined than that assumed in the current
version of the method.
Acknowledgments This work has been carried out within the S.A.G.G.I. research project, funded by the
Italian Ministry for the University and the Research (MUR) and led by Autostrade per lItalia S.p.A.

References
Antoniou S, Pinho R (2004a) Development and verification of a displacement-based adaptive pushover procedure. J Earthq Eng 8(5):643661
Antoniou S, Pinho R (2004b) Advantages and limitations of adaptive and non-adaptive force-based pushover
procedures. J Earthq Eng 8(4):497522
ASPI (AutoStrade Per lItalia S.p.a) (1992) S.A.M.O.A.: Survey, Auscultation and Maintenance of Bridges,
Autostrade per lItalia S.p.a., Rome
ATC (Applied Technology Council) (1996) Seismic evaluation and retrofit of concrete buildings. ATC-40,
RedwoodCity
ATC (Applied Technology Council) (2005) Improvement of nonlinear static seismic analysis procedures. Rep.
No. FEMA-440, Washington
Aydinoglu NM (2003) An incremental response spectrum analysis procedure based on inelastic spectral displacements for multi-mode seismic performance evaluation. Bull Earthq Eng 1:336 (Kluwer)
Aydinoglu MN, nem G (2007) Nonlinear performance assessment of bridges with incremental response
spectrum analysis (IRSA). In: ECCOMAS thematic conference on computational methods in structural
dynamics and earthquake engineering, Rethymno
Basz N, Mander JB (1999) Enhancement of the lifeline transportation module in HAZUS. Report No. Draft
#7, National Institute of Building Sciences, Washington
Biggs JM (1964) Structural dynamics. McGraw-Hill, New York
Caltrans (California Dept. of Transportation) (2006) Seismic design criteria ver. 1.4, Sacramento
Cardone D, Dolce M, Rivelli M (2008) Evaluation of reduction factors for high-damping design response
spectra. Bull Earthq Eng 9(1):273291 (Springer, Dordrecht)
Casarotti C, Pinho R (2007) An adaptive capacity spectrum method for assessment of bridges subjected to
earthquake action. Bull Earthq Eng (Springer) 5:377390
Casarotti C, Pinho R, Calvi GM (2005) Adaptive pushover based methods for seismic assessment and design
of bridge structures. ROSE Research Report No. 2005/06, IUSS Press, Pavia
CEN (Comit Europeen de Normalization) (1998) Eurocode 8: Design of structures for earthquake resistancepart 1: general rules, seismic actions and rules for buildings. PrEN 1998-1: 1998, CEN, Brussels
CEN (Comit Europeen de Normalization) (2004) Eurocode 8: Design of structures for earthquake resistancepart 2: bridges. PrEN 1998-2: 2004, CEN, Brussels

123

Bull Earthquake Eng (2011) 9:14631498

1497

CEN (Comit Europeen de Normalization) (2005) Eurocode 8: Design of structures for earthquake resistancepart 1: general rules, seismic actions and rules for buildings, PrEN 1998-1: 2005, CEN, Brussels
Chopra AK, Goel RK (2002) A modal pushover analysis procedure for estimating seismic demands for buildings. Earthq Eng Struct Dyn 31:561582
Clough RW, Penzien J (1975) Dynamics of structures. MacGraw-Hill, New York
Computers and Structures Inc (1999) SAP2000: Three dimensional static and dynamic finite element analysis
and design of structures. Computers and Structures Inc., Berkeley
DesRoches R, Leon RT, Dyke S (2003) Response modification of bridges. MAE Center Project ST-12
Dolce M, Cardone, D Croatto F (2005) Frictional behavior of steel-PTFE interfaces for seismic isolation. Bull
Earthq Eng 3(1):7599 (Springer, Dordrecht)
Dowell RK, Seible FS, Wilson EL (1998) Pivot hysteretic model for reinforced concrete members. ACI Struct
J 95:607617
Dutta A, Mander JB (1998) Seismic fragility analysis of highway bridges, INCEDE-MCEER center-to-center. In: Proceedings of the center-to-center workshop on earthquake engineering frontiers in transport
systems, Tokyo, pp 311325
Elnashai AS (2001) Advanced inelastic static (pushover) analysis for earthquake applications. Struct Eng
Mech 12(1):5169
Fajfar P (1999) Capacity spectrum method based on inelastic demand spectra. Earthq Eng Struct Dyn
28(9):979993
Isakovic T, Fischinger M (2006) Higher modes in simplified inelastic seismic analysis of single column bent
viaducts. Earthq Eng Struct Dyn 35:95114
Isakovic T, Fischinger M, Kante P (2003) Bridges: when a single mode seismic analysis is adequate?. Struct
Build 156:165173
Isakovic T, Lazaro MPN, Fischinger M (2008) Applicability of pushover methods for the seismic analysis of
single-column bent viaducts. Earthq Eng Struct Dyn 37:11851202
Jacobsen LS (1930) Steady forced vibrations as influenced by damping. Trans ASME 51:227
Kappos AJ, Paraskeva TS (2008) Nonlinear static analysis of bridges accounting for higher mode effects. In:
Workshop on nonlinear static methods for design/assessment of 3D structures, Lisbon
Kowalsky MJ (2002) A displacement-based approach for the seismic design of continuous concrete bridges.
Earthq Eng Struct Dyn 31:719747
Kowalsky MJ, Priestley MJN, MacRae GA (1995) Displacement-based design of RC bridge columns in seismic regions. Earthq Eng Struct Dyn 24(12):16231643
Mander JB, Priestley MJN, Park R (1988) Theoretical stress-strain model for confined concrete. J Struct Div
ASCE 114(8):18041826
Maroney BA, Romstad KM, Kutter B (1993) Experimental testing of laterally loaded large scale bridge abutments. In: Structural engineering in natural hazards mitigation: Proceedings of papers presented at the
structures congress93, vol 2. ASCE, New York, pp 10651070
Menegotto M, Pinto PE (1973) Method of analysis for cyclically loaded reinforced concrete plane frames
including changes in geometry and nonelastic behavior of elements under combined normal force and
bending. In: IABSE symposium on resistance and ultimate deformability of structures acted on by welldefined repeated loads, Zurich, pp 112123
Moschonas IF, Kappos AJ, Panetsos P, Papadopoulos V, Makarios T, Thanopoulos P (2009) Seismic fragility curves for greek bridges: methodology and case studies. Bull Earthq Eng 7:439468 (Springer,
Dordrecht)
Nielson B, DesRoches R (2007) Seismic fragility curves for typical highway bridge classes in the central and
southeastern United States. Earthq Spectra 23:615633
Paraskeva TS, Kappos AJ (2010) Further development of a multimodal pushover analysis procedure for seismic assessment of bridges. Earthq Eng Struct Dyn 39:211222
Paraskeva TS, Kappos AJ, Sextos AG (2006) Extension of modal pushover analysis to seismic assessment of
bridges. Earthq Eng Struct Dyn 35(11):12691293
Pinho R, Casarotti C, Antoniou S (2007) A comparison of single-run pushover analysis techniques for seismic
assesment of bridges. Earthq Eng Struct Dyn 36:13471362
Priestley MJN, Seible F, Calvi GM (1996) Seismic design and retrofit of bridges. Wiley, New York
Priestley MJN, Calvi GM, Kowalski M (2007) Displacement based seismic design of structures. IUSS Press,
Pavia
Reinhorn AM (1997) Inelastic analysis techniques in seismic evaluations. In: Krawinkler H, Fajfar P (eds)
Seismic design methodologies for the next generation of codes. Balkema, Rotterdam pp 277287
Saiidi M, Sozen M (1981) Simple nonlinear seismic analysis of R/C structures. J Struct Div ASCE
107(ST5):937952

123

1498

Bull Earthquake Eng (2011) 9:14631498

SeismoSoft (2006) SeismoStructa computer program for static and dynamic nonlinear analysis of framed
structures, freely downloaded from http://www.seismosoft.com
Takeda T, Sozen MA, Nielsen NN (1970) Reinforced concrete response to simulated earthquakes. J Struct
Eng Div ASCE 96(12):22572273

123

Вам также может понравиться