You are on page 1of 24

Bull Earthquake Eng (2010) 8:1397–1420

DOI 10.1007/s10518-010-9189-3

ORIGINAL RESEARCH PAPER

Hollow bridge-pier properties for response spectrum


analysis

Fabio F. Taucer · Carlo Paulotto · Gustavo Ayala

Received: 18 February 2009 / Accepted: 2 May 2010 / Published online: 22 May 2010
© Springer Science+Business Media B.V. 2010

Abstract The present paper proposes equivalent stiffness and energy dissipation proper-
ties of reinforced concrete hollow bridge piers to be used in the context of response spectrum
performance based assessment and design. The work is carried out by performing parametric
numerical analysis using a 2D fibre model calibrated against experimental results and by
varying the longitudinal steel reinforcement ratio, height over width ratio, normalised axial
force, level of confinement and concrete class of a rectangular hollow section reinforced with
Tempcore B500C steel. The results of the analysis are given in the form of charts and closed
form expressions for the yield curvature and moment, ultimate ductility, post yielding stiff-
ness ratio and energy dissipated of the section, and are translated to the member level through
the plastic hinge length approach. Likewise, the parameters of a Takeda model derived from
the parametric analysis are given for use in nonlinear time history analysis.

Keywords Bridge pier hollow section · Performance based design · Equivalent stiffness ·
Equivalent damping · Plastic hinge length · Takeda model

Abbreviations
DBD Displacement based design
PsD Pseudo-dynamic
RC Reinforced concrete

F. F. Taucer (B)
European Laboratory for Structural Assessment, European Commission – Joint Research Centre,
Via E. Fermi 2749, 21027 Ispra (VA), Italy
e-mail: fabio.taucer@jrc.ec.europa.eu
URL: http://elsa.jrc.ec.europa.eu/

C. Paulotto
ACCIONA, Madrid, Spain

G. Ayala
Universidad Nacional Autónoma de México, México, D.F., México

123
1398 Bull Earthquake Eng (2010) 8:1397–1420

ELSA European Laboratory for Structural Assessment


JRC Joint Research Centre
MP Medium Pier
TP Tall Pier

1 Introduction

Current seismic evaluation and design tendencies for reinforced concrete bridges, in which
performances under design conditions need to be estimated, require, even in the most simpli-
fied methods, an accurate description of the stiffness and energy dissipation characteristics
of the piers forming the substructure. This description involves not only the use of sound
analytical techniques, but also their calibration against existing results from experimental
tests on large scale specimens.
Considering that the most widely used approach for the Displacement Based Design
(DBD) of bridges is based on the use of secant stiffness and equivalent viscous damping
of the piers, both evaluated at maximum pier displacement, in this report the stiffness and
energy dissipation characteristics, necessary to estimate these parameters, are obtained for
reinforced concrete hollow rectangular bridge piers. The work involves the use of a continu-
ous nonlinear model calibrated against experimental results to derive the moment-curvature
behaviour of the section, translated into force-displacement properties at the level of the pier
by means of the plastic hinge approach.
The analysis starts from the identification of the parameters that play a major role in deter-
mining the behaviour of a pier section and their ranges of variation determined on the basis
of current practice and on the prescriptions contained in the Eurocodes.
The moment-curvature envelope of a generic section is determined through nonlinear
finite element parametric analysis under monotonically increasing curvatures, approximated
with bilinear curves defined by the yield curvature and moment, ultimate curvature and post-
yielding stiffness ratio of the section, all summarised in a series of charts and closed form
expressions. The cyclic behaviour of the section is reproduced through nonlinear analysis
under increasing cyclic loading and expressed in terms of a dimensionless parameter related
to the energy dissipated per unit length in a cycle.
The properties derived at the section level are used to compute the force-displacement
envelope and the energy dissipation characteristics of the pier for a given level of ductility
of the section; the calculations are performed with expressions derived from experimental
results for the computation of the plastic hinge length.
The equivalent properties of the bridge, namely, equivalent stiffness and equivalent damp-
ing at maximum displacement, are calculated based on the concept of a substitute linear
structure as originally defined by Gulkan and Sozen in 1967.
Lastly, a Takeda model with parameters derived from the parametric analysis is proposed
for assessment purposes by means of nonlinear time history analysis using the plastic hinge
approach.

2 Selection of experimental data

In spite of the wide use in the past decades of reinforced concrete (RC) hollow sections
for the construction of large bridge piers in seismic prone areas, there is only a limited
number of experimental tests describing the cyclic load-deformation and energy dissipation

123
Bull Earthquake Eng (2010) 8:1397–1420 1399

characteristics of such members. One of the most comprehensive experiments performed on


large scale bridge piers subjected to earthquake loading has been carried out by Pinto et al.
(1996) at the European Laboratory for Structural Assessment (ELSA) of the Joint Research
Centre (JRC) at Ispra, Italy; these tests were selected for calibrating the numerical models
used in the present research. The tests carried out at the JRC were performed on large-scale
models of reinforced concrete rectangular hollow sections, considering the combined effect
of axial loads and cyclic earthquake loading on piers designed according to EN 1998-1 (2004)
and EN 1998-2 (2004), and scaled to 1:2.5.
The models of the JRC were detailed to represent as best as possible the behaviour of the
prototypes, following a scaling criteria similar to that adopted by Stone and Cheok (1989),
whom after conducting a series of cyclic loading tests on full-scale and reduced scale circular
columns found that scale effects are not relevant when reinforcement details, including bar
diameter and vertical hoop spacing, are precisely scaled; Hoshikuma et al. (2001) arrived
to the same conclusion after performing tests on a full-scale square column and a 1/4 scale
replica model.
The experimental tests performed in Pinto et al. (1996) consisted in testing single piers
subjected to cyclic loading, as well as testing of a complete bridge structure by means of the
pseudo-dynamic (PsD) sub-structured test method, which integrates physical testing of the
pier models for which the load-displacement behaviour is to be determined, with numerical
models of those parts of the bridge for which the force-displacement response is known or

φ5 @50
F
330
136

φ5 @50 φ5 @50
346 346
136

136

14 φ14
φ5 @50
136
160

SLICES #
492

φ5 @50
8 0.24
6 φ12
5.60 7 0.24
492

6 0.24
LVDTs
LVDTs

136 5 0.24
492

4 0.24
1600

1280

20 φ8
136 3 0.24
φ5 @50
2 0.12
1 0.14

1.60
160

0.80 0.16
160 480 160
800

units [mm] units [m]


(a) (b)
Fig. 1 Medium pier of bridge B232 tested in Pinto et al. (1996) a Reinforcement layout of the pier section;
b Geometric characteristics of the pier and instrumentation

123
1400 Bull Earthquake Eng (2010) 8:1397–1420

Table 1 Mechanical properties of steel reinforcement1 (average values) for the medium pier of bridge B232
tested in Pinto et al. (1996)

Nominal bar Average bar f y ( MPa) f t ( MPa) εsu [%]


diameter ( mm) diameter ( mm)

φ52 4.7 699.5 730.9 1.6


φ8 8.2 503.4 563.0 12.3
φ 12 11.9 558.2 646.8 12.8
φ 14 13.8 477.2 577.7 13.0
1 Tempcore B500B steel reinforcement
2 Nominal bar diameter of the stirrups

Table 2 Mechanical properties


Cubic compressive Tensile Initial tangent
of concrete (average values) for
strength ( MPa) strength ( MPa) modulus (GPa)
the medium pier of bridge B232
tested in Pinto et al. (1996)
35.4 3.1 29.4

remains elastic (i.e., the deck). Several bridge configurations composed of three piers were
studied, according to their degree of irregularity in terms of the heights and capacities of the
piers. The results obtained from the medium pier of the bridge B232 configuration were used
for the calibration of the numerical model. The geometric characteristics and the reinforcing
steel lay-out of the pier model are presented in Fig. 1a; the mechanical characteristics of steel
and concrete are given in Tables 1 and 2, respectively, and were determined from tests on
a set of specimens from the construction of the model. The footing of the pier was rigidly
attached to the strong floor of the laboratory by means of post-tensioned steel bars passing
through the floor; a stiff steel cap was connected with bolts and epoxy resin to the top of
the pier for horizontal load application and for imposing the vertical loads (normalised axial
force equal to 0.1) needed to simulate the weight of the bridge superstructure.
The evaluation of the moment-curvature behaviour along the pier used in the calibration
of the numerical model of the section was carried out on the basis of data given by a set
of displacements transducers (LVDT’s) placed along two external opposite faces of the pier
model (Fig. 1b). The corresponding bending moment was calculated at mid-height of the
slice defined by the transducers, using the recorded values of the horizontal force and axial
load and the relative horizontal deflection to account for P- effects.

3 Numerical model

The parametric analysis described in the next section was carried out with a 2D fibre model
implemented in the finite element program CAST3M (CEA 2007) and calibrated against the
experimental results described in the previous section. Four different material models: rein-
forcement steel in the longitudinal direction, unconfined concrete, flange confined concrete
and web confined concrete, were used at the fibre level to build the section model.
The monotonic behaviour of the steel fibres is represented by a three-stage stress-strain
curve: linear elastic followed by a yielding plateau and a hardening zone modelled with a
fourth degree polynomial. The model for cyclic loading follows the monotonic curve until
the strain falls below a pre-established level after unloading from a postyielding position,

123
Bull Earthquake Eng (2010) 8:1397–1420 1401

Fig. 2 Stress-strain monotonic


relationship for confined concrete
in compression

from this point on, the stress-strain curve follows the Menegotto-Pinto model (1973); the
rebar buckling is modelled after Monti and Nuti (1992). The ultimate tensile strain of steel
was reduced by 30%, recognizing that under cyclic loading the ultimate tensile strain is in
general smaller than that obtained from monotonic testing (Priestley et al. 1996). The onset
of strain hardening was assumed at a strain equal to 0.02.
The monotonic compressive behaviour of the concrete fibres was modelled considering a
two branches law of Hognestad type, as shown in Fig. 2. The first branch, a parabolic func-
tion, defines the ascending part of the curve and goes from zero to the maximum compression
stress defined by the point (εc,c , f c,c ). The second branch, a descending straight line, rep-
resents the concrete softening behaviour after maximum strength until failure. The slope of
the second branch, equal to the product Z f c,c , and the value of the maximum compression
point, depend on the degree of confinement of the concrete:

f c, c = β f c (1)
εc, c = β εc
2
(2)
β − 0.85
Z =   (3)
β 0.1αw ωw + 0.0035 + εc,c

where β is a confinement parameter, f c is the maximum compressive stress of the uncon-


fined concrete, εc is the strain corresponding to f c and equal to 0.002, αw is a coefficient
that expresses the effect of both the longitudinal bars and the density of the stirrups on the
degree of confinement of the concrete core, and ωw is the mechanical volumetric ratio of the
stirrups, equal to:

Ast f yw (lw /s)
ωw = (4)
b0 h 0 f c
where Ast and lw are the area of the cross section and the length of hoops or ties in the
direction of confinement, f yw is the yielding stress of the stirrups, s is the distance between
stirrups along the member axis, and b0 and h 0 are the dimensions of the confined concrete
core measured from the centre-line of the stirrups. For unconfined concrete, the values of β
and Z are set equal to 1 and 100, respectively.
The values of the parameters β and αw were determined from the model proposed by
Mander et al. (1988), which has been accepted in EN 1998-2 (2004), Annex E, and can be
applied to all section shapes at all levels of confinement by defining a confinement parameter
λc equal to:

123
1402 Bull Earthquake Eng (2010) 8:1397–1420


σe 2σe
λc = 2.254 1 + 7.94 − − 1.254 (5)
fc fc
which is function of the effective confinement pressure:

σe = αw ρw f yw (6)

and

αw = αn αs (7)
 bi2
αn = 1 − (8)
6 b0 h 0
n
  
s s
αs = 1 − 1− (9)
2 · b0 2 · h0
Asw
ρw = (10)
sb
where ρw is the shear reinforcement ratio, Asw is the total area of hoops or ties in the direction
of confinement, n is the number of longitudinal restrained bars, bi is the distance between
consecutive restrained bars, and b is the dimension of the concrete core perpendicular to the
direction of the confinement under consideration, measured to the outside of the perimeter
hoop. For rectangular sections, the confinement effects should be evaluated in two orthogonal
directions, say directions 2 and 3, parallel to the long and short sides defining the rectangular
section. When the values of ρw in these two directions are not equal, the effective confining
pressure may be estimated as:

σe = σe2 σe3 (11)

Where the confinement effects are substantially different in each of the two directions the
effective confinement pressure is controlled by the smaller pressure; in this case σe is com-
puted as the minimum between σe2 and σe3 .
The confinement parameter β for the confined concrete fibres was made equal to the con-
finement parameter λc given in Eq. (5); for the unconfined concrete, the value of β was set
equal to 1.
Moreover, for the confined concrete a third branch was considered after the softening
compression branch and before reaching failure: a zero slope straight line defining a com-
pression plateau. This additional condition accounts for the residual strength of the concrete
core for important axial, post-peak deformations. Following the recommendations of Park
et al. (1982), a residual strength equal to 20% of the peak strength was assumed in the model.
Unloading from the envelope follows a law similar to the one proposed by Mercer and
Martin (1987); no strength degradation is considered. A bilinear model was used to represent
the behaviour of concrete under monotonic tensile loading, unloading to zero stress after
reaching the maximum tensile stress. Details on the model for cyclic tensile loading can be
found in Guedes et al. (1994).
The calibration of the numerical model focused on a correct representation of the initial
stiffness of the section, considering the cracked and uncracked stiffness, and on predicting
the section behaviour after yielding. The results indicate that the numerical model is able to
follow the skeleton curve of the section, as shown in Fig. 3 for slice #1 (see Fig. 1b), while
giving a good estimation at different curvature amplitudes of the dissipated hysteretic energy
for the first four slices (Fig. 4), in spite of the more pronounced pinching of the numerical

123
Bull Earthquake Eng (2010) 8:1397–1420 1403

Fig. 3 Comparison between numerical and experimental results relative to slice #1 of the pier when subjected
to the entire duration of the design earthquake

Fig. 4 Comparison between the energy dissipated per cycle by the numerical model and by the sections
belonging to different pier slices for cyclic tests of increasing amplitude

model with respect to the experimental results. The mechanical properties of steel and con-
crete were equated to those obtained from the experimental test as given in Tables 1 and 2
considering a 30% reduction on the ultimate strain of steel εsu . The remaining parameters of
the numerical model are given in Table 3; full details concerning the calibration procedure
are given in (Paulotto et al. 2007).

4 Parametric analysis

The aim of the parametric analysis was to evaluate the moment-curvature behaviour, ductil-
ity capacity and energy dissipated of reinforced concrete pier sections of various geometric

123
1404 Bull Earthquake Eng (2010) 8:1397–1420

Table 3 Parameters used to compute the maximum stress point (εcc , f cc ) and slope of descending branch of
the confined concrete constitutive model in CAST3M
f c1 28.3 MPa εc 0.002
Asw2 157.1 mm2 Asw3 39.3 mm2
b02 779 mm b03 139 mm
h 02 139 mm h 03 779 mm
s2 60 mm f yw 699.5 MPa
lw 2860 mm Ast 19.6 mm2
σe2 1.361 MPa σe3 1.908 MPa
σe 1.612 MPa αn 0.768
ωw 3 0.214 αw 0.579
β = λc 1.348 Z 18.9
1 Concrete cylinder strength, computed as 0.8 times the cubic strength given in Table 2
2 Stirrup spacing measured in the mock-up at slice #1
3 The mechanical volumetric ratio of stirrups is computed using b and h
02 02

configurations and reinforcement layouts subjected to different levels of axial forces. The
following parameters that play a major role in determining the behaviour of the section were
considered: concrete class and steel reinforcement yield strength, wall thickness, section
aspect ratio, longitudinal reinforcement ratio, axial load level, and confinement level. The
definition of the range of variation for each of these parameters is discussed in the following
paragraphs, while the values that were considered in the analysis are reported in Table 4.
For each of the 2700 combinations of these values, two nonlinear static analyses, one mono-
tonic, and one cyclic, were performed in order to obtain the skeleton curve and the damping
properties of the corresponding pier section. Finally, each skeleton curve was approximated
through a bilinear curve.
Tempcore B500C reinforcing steel belonging to class C as defined by Normative Annex
C of EN 1992-1-1 (2004) was considered in the analysis, assuming a characteristic yield
strength f yk of 500 MPa and characteristic tensile strength f tk equal to 1.19 f yk (Priestley
et al. 1996), elongation εsu at maximum force equal to 0.11 (monotonic) and strain at the onset
of hardening equal 0.02. According to EN 1998-1 (2004), article 7.2.1 (1), the prescribed
concrete class in plastic regions should not be lower than C20/25, and not higher than C40/50.
In the parametric analysis, concrete classes C25, C30 and C35 were considered. Following
the provisions of EN 1998-1 (2004), article 4.3.3.4 (4), the element properties derived in the
parametric analysis were computed from the mean values of the material properties based
on information provided in Table 6.1 of EN 1992-1-1 (2004) and in Annex E of EN 1998-2

Table 4 Range of values used


Wall thickness (m) 0.40
in the parametric analysis
Concrete C25 C30 C35
H/B 1.0 1.5 2.0 2.5 3.0
Steel Tempcore B500C
ρL 0.005 0.010 0.020 0.030 0.040
νk 0.10 0.20 0.30 0.40
λc 1.0 1.2 1.3 1.4 1.6 1.8 2.0

123
Bull Earthquake Eng (2010) 8:1397–1420 1405

(2004). The mean values of the compressive stress f cm and yield stress f ym of concrete and
steel were calculated as 8 + f ck ( MPa) and 1.15 f yk , respectively, with f ck and f yk equal to
their characteristic values.
From a survey of a number of bridge designs with hollow sections it was observed that the
thickness t of the walls varies between 0.30 and 0.50 m; in the parametric analysis a constant
value of t equal to 0.40 m was chosen for the wall thickness . A monocellular hollow pier with
the outer dimension B of the flange constant and equal to 2.00 m was considered, resulting
in a width to thickness ratio of the flange B/t constant and equal to 5, and in a section aspect
ratio of 2.0 at the maximum wall slenderness ratio allowed by article 6.2.4 (2) of EN 1998-2
(2004) and equal to 8. Aspect ratios H/B ranging between 1.0 and 3.0 were considered in
the parametric analysis, exceeding the maximum allowed slenderness ratio for aspect ratios
larger than 2.0.
The longitudinal reinforcement ratio ρ L , equal to As /Ac , where As and Ac are the total
areas of longitudinal reinforcement and concrete cross-section, was varied between 0.005
and the maximum allowed by article 5.4.3.2.2 (1) of EN 1998-1 (2004) and equal to 0.04;
the steel reinforcement was distributed in two layers and uniformly distributed across the
section.
The normalized axial force, νk , was varied between 0.1 and 0.4, a range commonly found
in design practice, and is equal to:
N Ed
νk = (12)
Ac f ck
where N Ed is the design axial force corresponding to the seismic condition, and f ck is the
characteristic strength of concrete, equal to f cm − 8 (units in MPa).
The variation of the level of confinement of the cross section was established by evaluating
Eqs. (5) to (10) (with f c equal to f cm ) considering the possible combinations of transverse
reinforcement layout as a function of the diameter of the longitudinal steel, the longitudinal
steel reinforcement ratio, the normalised axial force and the restrictions imposed by EN 1998-
2 (2004) on rebar spacing. For longitudinal steel reinforcement diameters varying between
16 and 32 millimetres at 100, 150 and 200 millimetre spacing, the transverse reinforcement
ratio ρw was found to vary from 0.003 to 0.018, leading to confinement levels λc varying
between 1.0 and 2.0 (Paulotto et al. 2007).
The failure of the section was calculated when the reinforcing steel attains its ultimate
strain in tension or compression, when the confined concrete reaches its ultimate compres-
sive strain, or when the strength of the section decreases to 80% of its maximum value. The
ultimate tensile strain of steel was set at 70% of the strain εsu at maximum stress under
monotonic load, recognizing the reduced deformation capacity of steel under cyclic loading
(Priestley et al. 1996). The ultimate compressive strain of the confined concrete was computed
according to EN 1998-2 (2004), article E.2.1:

1.4ρs f ym εsu
εcu,c = 0.004 + (13)
f cm,c
where ρs = 2ρw for orthogonal hoops, εsu∗ is the elongation at maximum stress of the rein-

forcement steel under cyclic loading, made equal to 0.70εsu , and f cm,c is the mean value
of the compressive strength of the confined concrete. Equation (13) has been formulated
from considerations on confined sections under axial compression: when used to estimate
the ultimate compression strains of sections subjected to bending, or combined bending
and axial compression, Eq. (13) tends to be conservative by at least 50% (Priestley et al.
1996). Mean values of the concrete strength, f cm , are used in Eq. (13), as well as in the

123
1406 Bull Earthquake Eng (2010) 8:1397–1420

Fig. 5 Construction of the


moment-curvature bilinear
approximation of the nonlinear
skeleton curve of the pier section

estimation of the level of confinement and of the properties of the constitutive model of
the concrete fibres following the provisions of Annex E of EN 1998-2 (2004), which states
that mean values should be used for the estimation of the deformation capacity of plastic
hinges.
The nonlinear skeleton curves obtained from the parametric analysis were approximated
through bilinear curves, based on first yield, when the first steel fibre reaches the yield strain
in tension or when the first extreme fibre in compression attains a strain of 0.002, and on
failure of the section, as described in the previous paragraph. The line that joins the origin
and the first yield point gives the initial slope of the bilinear curve, whereas the line that
extends through the failure point and balances the areas between the actual and the idealized
moment-curvature relationships beyond the first yield point gives the slope of the second
branch (Fig. 5).
It is worth noting that sections with different detailing, and hence with different rein-
forcement ratios ρw , may lead to the same value of the confinement parameter λc . Since the
ultimate compressive strain εcu,c depends on the amount of ρw , for the same value of λc ,
different values of εcu,c , and hence of maximum curvature, may be obtained.

5 Equivalent section properties

The moment-curvature capacity curve of a generic section is represented through a bilin-


ear relationship, as defined in the previous section, and four parameters: yield curvature
and moment, χ y and M y , and ultimate curvature and moment at failure of the section, χu
and Mu , which were used to summarize the results of the parametric analysis in a series
of charts (Paulotto et al. 2007). An example of these charts is shown in Fig. 6 for ρ L
equal to 0.02, where the results are expressed in terms of the following dimensionless
parameters:

χy H (14)
χu
μu = (15)
χy
My
(16)
f cm B H 2
Mu −M y
χu −χ y
α= My
(17)
χy

123
Bull Earthquake Eng (2010) 8:1397–1420 1407

Fig. 6 Results from the parametric analysis ( f cm = 33 MPa, ρ L = 0.020, λc = 1.2)

By fitting the numerical results, analytical expressions were determined to estimate the yield
curvature and moment:
1
χ y = 0.00971ρ L0.114 [rad] (18)
H

−3 (ρ L +0.063)
My H
= 140 C1 C2 (ρ L + 0.013) [N, mm]
BH2 B

C1 = 1650 (0.04 − ρ L )2 + 0.45 ; C2 = νk + (28ρ L + 0.31)2 − 0.07 (19)

Equations (18) and (19) correspond to the average values obtained for the confine-
ment parameter λc varying between 1.2 and 2.0, over which the secant-to-yield stiff-
ness remains practically constant with maximum variations of the yield point (χ y , M y )
between 2 and 9% for ρ L equal to 0.005 and 0.04, respectively, with respect to the aver-
age obtained for λc equal to 1.6. Equation (19) is given in dimensional form, as M y
is not influenced by the concrete classes considered in the analysis (i.e., C25 through
C35).
By rewriting the product 0.00971ρ L0.114 as υε y it is possible to compare Eq. (18) with
similar expressions proposed in literature for the estimation of the yield curvature. For ε y
equal to 0.0025 and for ρ L varying between 0.005 and 0.04, υ varies in Eq. (18) from 2.12
to 2.69, with a variation of 12% around its mean value of 2.41. Such values are consistent

123
1408 Bull Earthquake Eng (2010) 8:1397–1420

with those proposed by the International Federation for Structural Concrete (2006) that set
υ equal to 2.25 within an error of 10% for rectangular hollow sections. Other expressions
proposed are by Priestley et al. (1996) that set υ equal to 2.14 ± 10% for rectangular sections,
and by Priestley and Kowalsky (1998) that set υ equal to 2 ± 5% for rectangular walls with
0.5% reinforcement ratio plus end reinforcement (up to 1.5%). The difference of υ between
the latter of these expressions and that proposed in Eq. (18) is due to the different section
shape, the amount of longitudinal reinforcement and the linearisation procedure considered
in deriving χ y .
Concerning the ratio of post-yielding to initial stiffness α, the charts in (Paulotto et al.
2007) indicate that for λc equal or larger than 1.2, α is always equal or lower than 0.01,
suggesting that the bilinear diagram may be taken as elastoplastic (i.e., α equal to zero). For
the particular case of λc = 1.2 and νk equal to 0.3 and 0.4, α is negative, however, since its
value is low, varying between 0 and −0.02, α is assumed to be equal to zero, which is a valid
approximation considering that the expressions proposed are intended to be used within the
framework of equivalent linear analysis. For λc equal to 1.0, larger values of α, as large as
0.18, are obtained, due to curve fitting of the numerical envelope with the bilinear diagram at
low levels of ductility. In practice, α may also be taken equal to zero for λc equal to 1.0, since
at low ductilities the difference between the yield and the ultimate points is not relevant. As
a result, Eqs. (18) and (19) are also applicable for λc equal to 1.0.
Concerning μu , some general considerations can be extrapolated: for λc ≤ 1.4, μu
decreases with the increase of H/B, ρ L and νk , while μu may be approximated as being
independent of the different concrete classes considered as failure of the section is controlled
by fracture of the steel fibres in tension. The values of μu obtained from the charts of Fig. 6
and given in (Paulotto et al. 2007), when used in conjunction with Eq. (18) should be recal-
culated by multiplying the value of μu from the chart by the ratio of the corresponding yield
curvature χ y obtained from the chart and the yield curvature given by Eq. (18). The derivation
of closed form expressions for the calculation of μu is presented in Sect. 7 within the context
of the proposal of a simplified design procedure.
For flange width to thickness ratios different from that considered in the parametric anal-
ysis (i.e., B/t equal to 5), the expression for computing M y given in Eq. (19) remains
unchanged, as the neutral axis falls within the flange in compression when computing the
ultimate moment of the section – equal to the yield moment for the particular case of elasto-
plastic behaviour; the yield curvature slightly decreases with the increase of the wall width,
however, for practical purposes the yield curvature may be considered independent of the
width and equal to the value given in Eq. (18). Likewise, the value of ρ L used in Eq. (19)
and in the chart of Fig. 6, for B/t ratios different from 5, must be computed by multiply-
ing the actual value of the steel longitudinal reinforcement ratio by a factor equal to 5t/B.
The normalised axial force νk and confinement parameter λc are computed using the actual
geometry and f cm of the cross section.
The hysteretic energy dissipated by the considered sections was evaluated from the para-
metric analysis after nonlinear analyses performed under increasing cyclic curvature; the
results are expressed in terms of a dimensionless parameter:
W
η= (20)
2π Mmax χmax
where W is the energy dissipated in one cycle, Mmax and χmax are the maximum moment
and curvature cyclic amplitude, respectively. The results indicate that η does not depend on
the section aspect ratio, while it depends strongly on the normalized axial force, although
this dependence becomes weaker as the longitudinal reinforcement ratio increases. It was

123
Bull Earthquake Eng (2010) 8:1397–1420 1409

also found that by increasing the confinement of the section, the section ductility increases
without any relevant changes in the dissipated energy; based on this result, all the cyclic
analyses were conducted assuming λc equal to 2.0. The results for sections in which C25
concrete is used are shown in Fig. 7, which may be approximated, for νk ≤ 0.30, by the
following expression in terms of the curvature ductility, μ:

νk − 0.1 1
η = 1− 0.96ρ L0.2 1 − √ (21)
78 · ρ L μ

Equation (21) may also be used for concrete classes C30 and C35, as well as for widths dif-
ferent from that considered in the analysis and within the ranges typically found in practice.
All the expressions presented in this section correspond to Tempcore B500C steel.

6 From section to member properties

Using the plastic hinge approach, the properties derived at the section level are used to com-
pute the force-displacement envelope and the energy dissipation characteristics of the piers,
from which the pier equivalent properties of stiffness and damping ratio, K eq and ξeq , are
evaluated.
The equivalent stiffness of a cantilever pier of height L is defined as the secant stiffness
K eq at displacement  at the top of the pier:

 =  y μd (22)
χy · L2
y = (23)
3  

3L p Lp
μd = 1 + (μ − 1) α + 1 − 0.5 (24)
L L

with  y and μd equal to the yield displacement and the displacement ductility of the
pier, respectively, with L p denoting the plastic hinge length. Considering that the force-
displacement envelope of the pier is bilinear, the equivalent stiffness is expressed as:

K eq = K y when μd ≤ 1
(25)
K eq = 1+αdμ(μd d −1) K y when μd > 1
3 · My
Ky = (26)
χy · L 3
1
αd =   (27)
3L p L
1+ αL 1 − 0.5 Lp

where K y and αd are the secant-to-yield stiffness and the post-yield stiffness ratio of the pier,
respectively.
The equivalent damping ratio of the pier ξeq is calculated according to Jacobsen (1930)
as a function of Wd and E d , equal to the energy dissipated and energy stored by the pier in
a cycle of maximum ductility μd :

123
1410 Bull Earthquake Eng (2010) 8:1397–1420

Fig. 7 Dimensionless energy dissipated by the pier section for cycles of different ductility

Wd
ξeq = (28)
2 · π · Ed

Wd = W · L p (29)
χy L
Ed = μd M y [1 + α (μ − 1)] (30)
3

123
Bull Earthquake Eng (2010) 8:1397–1420 1411

By substituting Eqs. (20), (29) and (30) into Eq. (28), and by making Mmax and χmax equal to
M y [1+α(μ−1)] and χ y μ, respectively, the following expression for the equivalent damping
ratio of the pier is obtained:
μ Lp
ξeq = 3η (31)
μd L
The expressions presented in this section give good estimates of the load-deformation
response of members where the relative contribution of shear with respect to flexural
deformation is not important, i.e., for members with shear span-to-depth ratios larger than
2.0 ∼ 2.5, as well as for members with low shear span-to-depth ratios where the contribution
of shear deformation is relevant, following the approach exposed in the following paragraphs.
According to the plastic hinge method, the displacement at the top of a cantilever pier of
length L is computed as the sum of the contribution of the deformations from a plastic and an
elastic region, defined as a function of the distribution of curvatures along the member. The
curvatures along the plastic region span over a length L p and are considered constant and
equal to the curvature at the critical section of the member, while the curvatures along the
elastic region decrease linearly to zero from the curvature at yield. The displacement  at the
top of the pier is then computed from the first moment of inertia of the curvature distribution
about the top of the pier:
 
χy L 2   Lp
= + χmax − χ y L p L − (32)
3 2
where χmax is the maximum curvature at the critical section of the pier (i.e., at the base).
Equation (32) considers that the behaviour of the pier is elasto-plastic (post-yield stiffness
ratio α of the section equal to zero), which is a valid approximation for bridge piers reinforced
with Tempcore B500C steel reinforcement and detailed to undergo plastic deformations (i.e.,
λc > 1). The length of the plastic hinge may be computed from the results of experimental
tests by solving Eq. (32), such that for the maximum curvature measured at the base of the pier
at different ductility levels, the corresponding maximum displacement measured at the top of
the pier is obtained; the yield curvature is assessed from the experimental moment-curvature
diagram of the section. Following this procedure and using the experimental results from
Pinto et al. (1996), the values of L p , presented graphically in Fig. 8 for the tall, medium and
short piers, were obtained as a function of the curvature ductility μ of the critical section, as
expressed by the following expression:
L
μ−1
L = c L μ L −1 with 1 ≤ μ ≤ μ L
p

Lp (33)
L = cL with μ ≥ μ L
where c L and μ L are two parameters that depend on the shear span-to-depth ratio L/H , and
are equal, for the medium (MP) and tall piers (TP) (L/H equal to 3.5 and 5.25), to 0.0624
and 9, and for the short pier (A1) (L/H equal to 1.75), to 0.127 and 5, respectively. The
experimental points given in Fig. 8 for the MP and TP piers for ductilities larger than 13
where not considered in deriving Eq. (33), as for these ductilites the piers were considered to
have failed after a reduction of more than 20% of their ultimate capacity. Since these results
were derived from a very limited number of tests, it would be desirable that a larger set of
experimental data is used to increase the reliability of the proposed curves.
An important feature of the procedure exposed above to calculate the length of the plastic
hinge is that it allows determining in a simplified manner the flexibility of a short pier, account-
ing in an empirical way for the contribution of shear deformations that otherwise would need

123
1412 Bull Earthquake Eng (2010) 8:1397–1420

Fig. 8 Values of the ratio


between the plastic hinge length
and the pier height as a function
of the ductility level of the critical
section of the pier and the pier
aspect ratio

Fig. 9 Equivalent stiffness in terms of pier ductility: comparison between numerical and experimental values
for the short (A1), medium (MP) and tall (TP) piers

to be obtained from more refined and computationally expensive analytical methods: this is
achieved by deriving the plastic hinge length of the short pier so that both flexural and shear
deformations are included.
As an example, the proposed procedure is applied to derive the equivalent stiffness (Fig. 9)
and equivalent damping ratio (Fig. 10) of the A1 pier tested in Pinto et al. (1995) and
of the medium and tall piers (MP, TP) of the B231C bridge tested in Pinto et al. (1996).
The piers were constructed using Tempcore B500B steel rebars—with properties within the
range of class C steel reinforcement for the longitudinal bars of the flanges, thus making
the results comparable with the expressions proposed in Sects. 5 and 6—and C25 concrete
∗ = 33 MPa), with an aspect ratio H/B of the section equal to 2.0 (H and B equal to 1.6
( f cm
and 0.8 m, respectively) and a normalised axial force νk equal to 0.10. In Table 5 are listed
the remaining characteristics of the piers, showing the values of the parameters used to derive
the equivalent properties and the bilinear envelopes at the section and pier level; the length
of the plastic hinge was derived from the curves shown in Fig. 8.

123
Bull Earthquake Eng (2010) 8:1397–1420 1413

Fig. 10 Equivalent damping in terms of pier ductility: comparison between numerical and experimental
values for the short (A1), medium (MP) and tall (TP) piers

Table 5 Parameters for deriving


Equation A1 pier MP pier TP pier
the bilinear force displacement
envelopes and equivalent
L (m) 2.8 5.6 8.4
stiffness K eq and damping ratio
ξeq of the short (A1), medium ρL 0.009 0.012 0.012
(MP) and tall (TP) piers λc 1.24 1.22 1.22
ρw 0.0033 0.0033 0.0033
χ y (mrad) (18) 3.55 3.67 3.67
M y (kN · m) (19) 3823 4783 4783
α 0 0 0
μ1u From charts 16.0 15.9 15.9
η (21) 0.28 0.30 0.30
 y (m) (23) 0.00927 0.0383 0.0862
L p (m) (33) 0.36 0.35 0.52
1 μ is calculated as μ from αd (27) 0 0 0
u u
chart multiplied by χ y from chart K y (MN/m) (26) 147 22.3 6.61
divided by χ y from Eq. (16)
2 μ is given by Eq. (22) by μ2du (24) 6.4 3.7 3.7
du
setting μ = μu 3 (MN/m)
K eq (25) 23.2 6.01 1.78
3 K and ξ calculated at
eq eq 3
ξeq (31) 0.27 0.24 0.24
μd = μdu and μ = μu

7 Ultimate deformation

The displacement capacity u of a cantilever pier may be computed from Eqs. (18), (32)
and (33), by substituting  with u , χmax with χu , equal to χ y μu , and μ by μu ; the ultimate
curvature ductility μu being obtained from the charts generated through the parametric anal-
ysis (see Fig. 6) as explained in Sect. 5. In the following paragraphs a simplified procedure
to design the cross section of a bridge pier to meet a given target performance is presented,
based on closed form expressions derived for the ultimate ductility of the section.

123
1414 Bull Earthquake Eng (2010) 8:1397–1420

Fig. 11 Schematic variation of


μu with respect to λc according
to Eq. (34) for a given set of
H/B, νk and ρ L values

Fig. 12 Variation of μu,min and μu,max in terms of νk and ρ L

Insight on the variation of the ultimate curvature ductility with respect to the parameters
considered in the parametric analysis may be obtained by examining the variation of μu with
respect to λc for a given set of H/B, νk and ρ L values, as shown schematically in Fig. 11,
where μu is approximated as varying linearly from a minimum value equal to μu,min at λc
equal to 1, up to a maximum value equal to μu,max at λc equal to λ∗c , and from there on
remaining constant, as expressed by the following expression:
 
μ −μu,min
μu = μu,min + u,max λ∗c −1 (λc − 1) for 1 ≤ λc < λ∗c
(34)
μu = μu,max for λc ≥ λ∗c

Figure 11 reflects that failure of the section in the rising branch of the curve (for λc < λ∗c ) is
determined by the ultimate compressive strain of concrete, while in the constant part of the
curve (for λc ≥ λ∗c ) failure of the section is determined by fracture of the steel reinforcement
in tension.
With the purpose of simplifying the design process, the ultimate curvature ductility is
averaged over the considered range of H/B values, with maximum variations of 15% and
5% for μu,min and μu,max , respectively.

123
Bull Earthquake Eng (2010) 8:1397–1420 1415

Fig. 13 Variation of λ∗c in terms of νk and ρ L

The variation of μu,min , μu,max , and λ∗c with respect to νk is plotted in Figs. 12 and 13 for
different values of ρ L , leading to the following approximated closed form expressions:

− (20ρ L −0.85)2 +0.30
μu,min = 1.06 νk ; μu,max = 6.33 (νk + 2.37) (35)
λ∗c = G 1 νk + G 2 (36)

where G 1 and G 2 are parameters equal to (1.28; 1.44; 1.44; 1.04; 0.92) and (1.10; 1.12;
1.30; 1.52; 1.64) for ρ L equal to (0.005; 0.01; 0.02; 0.03; 0.04), respectively. Equations (35)
and (36) reflect the variation of the ultimate curvature ductility of the section resulting from
balancing the compression and tension forces of concrete and steel as a function of the levels
of νk and ρ L . In particular, larger values of μu,max are achieved at higher levels of νk , since
the increased axial compression allows for larger rotations of the section while maintaining
the strain associated to fracture of the steel reinforcement in tension.
Equations (34), (35) and (36) indicate that at low levels of νk and ρ L it is sufficient to pro-
vide a limited amount of λc to reach μu,max , while at high levels of νk and ρ L high values of λc
are necessary to develop μu,max . This information may be used for the assessment and design
of a bridge pier in relation to a given target performance, which may be defined, in terms of
a maximum displacement ductility μd , a maximum drift , or a maximum displacement 
of the pier.
If the maximum displacement ductility of the pier is used as target performance, the vari-
ation of μ as a function of μd may be expressed by rearranging Eq. (32) and by substituting
 with  y μd , with  y = χ y L 3 /3, so that the following expression is obtained:

1
μ=1+   [μd − 1] (37)
L L
3 Lp 1 − 0.5 Lp

where L p /L is given by Eq. (33). Note that for μ < μ L , L p /L is a function of μ, so that
Eq. (37) needs to be solved iteratively for μ. The plot of Eq. (37) for μd equal to 2, 3 and 4
is given in the chart of Fig. 14 for the medium and tall piers for νk equal to 0.1. The chart
suggests that to reach a given level of μd high levels of λc are required at large values of ρ L .

123
1416 Bull Earthquake Eng (2010) 8:1397–1420

Fig. 14 Proposed chart for design and assessment of MP and TP pier sections with νk = 0.10 using displace-
ment ductility μd as target performance

Likewise, if the maximum drift δ of the pier is used as target performance, the variation
of μ with respect to  is obtained by expressing μd as:

δL
μd = (38)
y

so that by substituting Eqs. (23) and (18) into Eq. (38), which is then substituted into Eq. (37),
the following expression is obtained:

 
1 3 δH
μ=1+   −1 (39)
3
Lp L
1 − 0.5 Lp 0.00971ρ L0.114 L
L

In the same way as for Eq. (37), Eq. (30) needs to be solved iteratively for μ when μ is
less than μ L . The plot of Eq. (39) as a function of λc for the non-dimensional values δ H/L
corresponding to 0.005, 0.006 and 0.007 is given in Fig. 15 for the medium and tall piers;
the intersection between Eqs. (39) and (34) provides the performance point of the pier.
The use of these charts is illustrated in the following example for a section with νk equal
to 0.1. For assessment purposes of the maximum displacement capacity of the pier, suppose
the section is detailed with ρ L equal to 0.02 and λc equal to 1.3, the charts indicates that the
maximum displacement ductility and maximum drift that the pier can develop is equal to 2.9
and 0.0061L/H , respectively, both associated to a value of μu equal to 11.7. Likewise, for
design, suppose that the target displacement ductility is equal to 3 or that the target drift is
equal to 0.007L/H , with the constraint of developing a minimum flexural strength corre-
sponding to ρ L equal to 0.03, the charts indicates that the section should be confined with a
detailing corresponding to λc equal to 1.46 and 1.51, respectively.
For the case where the maximum displacement  is used as target performance, it is
sufficient to substitute into Eq. (39) and in Fig. 15 the term δ H/L by δ H/L 2 .

123
Bull Earthquake Eng (2010) 8:1397–1420 1417

Fig. 15 Proposed chart for design and assessment of MP and TP pier sections with νk = 0.10 using δ H/L
as target performance

8 Takeda model

The parameters of a Takeda model to be used in the context of nonlinear time history anal-
ysis to assess the earthquake response of a bridge structure with rectangular RC hollow
piers are given in the following paragraphs, assuming the following properties are known:
H, B, L , νk , ρ L and λc . The Takeda Model is defined by a moment-rotation envelope and
by two factors, an unloading stiffness parameter a and a reloading stiffness parameter b, that
determine the rules for cyclic loading and unloading.
The moment-rotation envelope is bilinear and is defined by the yield rotation and moment,
θ y = χ y L p and M y , as given by Eqs. (18), (19) and (33), and by the ratio of the post-yielding
to initial stiffness r, equivalent to α of Eq. (17) and set equal to zero. The parameters a
and b are chosen such that the energy ξT dissipated by the Takeda model, as given by the
expression proposed by Loeding et al. (1988) in Eq. (40), is equal to the energy dissipated η
by the section as obtained from the parametric analysis.
 
 

2 3 1 r bμ 1 1
ξ T = ξ0 + 1 − μa−1 − 1− +1 2−b· 1−
π 4 4 γ μ μ
   
2
1 r b2 μ 1
−μa−1 γ − 1− (40)
4 γ μ
Where γ = r μ − r + 1, a is the unloading stiffness factor and b is the reloading stiffness
factor. The values of a and b are calculated for ξ0 equal to zero, i.e., only the contribution of
hysteretic damping is taken into account.
Since different combinations of a and b may yield the same value of damping, it was
decided to fix one of the parameters while varying the other. In view of the fact that the reload-
ing stiffness parameter for bridge piers tends to be near zero (i.e., ‘thin’ cycles) (Blandon
and Priestley 2005), it was decided to keep b constant, while varying parameter a. It was
found that a value of b equal to 0.3 was the only value that allowed fitting Eq. (40) with
the energy dissipated by the numerical model as given from the parametric analysis, while
varying parameter a between 0.01 and 0.93, as shown in Fig. 16 for ρ L equal to 0.02. The

123
1418 Bull Earthquake Eng (2010) 8:1397–1420

Fig. 16 Comparison of the hysteretic dissipated energy at different ductility levels between test results and
Takeda model with b = 0.3 and a computed from Eq. (41) for different levels of νk

Table 6 Takeda model: a values


νk ρL
as a function of ρ L and νk , with
b = 0.3 and r = 0 0.005 0.01 0.02 0.03 0.04

0.1 0.62 0.47 0.28 0.13 0.01


0.2 0.79 0.61 0.41 0.26 0.13
0.3 0.88 0.75 0.54 0.39 0.27
0.4 0.93 0.82 0.64 0.47 0.33

values of a are given in Table 6, and are fitted by the following closed-form expression as a
function of the amount of longitudinal reinforcement ρ L and axial load ratio νk :

a = 1.1 (νk − 1) − 7.2 (ρ L − 0.005)0.73 + 0.64 (41)

9 Conclusion

The present paper provides expressions to determine the equivalent secant stiffness and
equivalent damping of reinforced concrete rectangular hollow sections of bridge piers of var-
ious geometries and steel layouts reinforced with Tempcore B500C steel at different levels
of normalised axial force to be used within the framework of displacement based assessment
and design. The expressions were derived from moment-curvature monotonic and cyclic
curves at the pier section obtained from parametric analysis using a fibre model calibrated
against experimental results performed on large scale specimens.
The results at the section level are expressed in charts and closed-form equations based on
bilinear moment-curvature envelops and equivalent damping representing the energy con-
tained in the curves obtained from the parametric analysis. The results at the section level
are translated to the pier level following the plastic hinge approach; expressions calibrated
against experimental results for determining the length of the plastic hinge length in terms
of ductility are proposed for tall to medium, and short piers.

123
Bull Earthquake Eng (2010) 8:1397–1420 1419

A series of charts for assessing and designing the performance of a bridge pier in terms
of target ductility, drift and displacement are proposed, based on the results of the ultimate
ductility obtained at the section level from the parametric analysis.
Lastly, the parameters of a Takeda model fitting the moment-rotation envelopes and equiv-
alent damping from the parametric analysis are proposed for the assessment of the perfor-
mance of a bridge structure using nonlinear analysis with plastic hinges at the base of the
pier.
It is believed that the simplified procedures and expressions proposed will provide a useful
tool for the preliminary assessment and design of bridges, as well as for the optimization of
cross-section properties and detailing; complex structures (i.e., irregular bridges) should be
designed/checked on the basis of adequate analysis methods and models.

Acknowledgements The technical support given by Pierre Pegon and Artur Pinto of the Joint Research
Centre is greatly appreciated.

References

Blandon CA, Priestley MJN (2005) Equivalent viscous damping equations for direct displacement based
design. J Earthq Eng 9(2):257–278
CEA (2007) In: Cast3M. French Atomic Energy Commissariat. http://www.cast3m.cea.fr. Accessed 15 July
2007
EN 1992-1-1 (2004) Eurocode 2: design of concrete structures—Part 1-1: general rules and rules for buildings.
European Committee for Standardization, Brussels, Belgium
EN 1998-1 (2004) Eurocode 8: design of structures for earthquake resistance—Part 1: general rules, seismic
actions and rules for buildings. European Committee for Standardization, Brussels, Belgium
EN 1998-2 (2004) Eurocode 8: design of structures for earthquake resistance—Part 2: bridges. European
Committee for Standardization, Brussels, Belgium
Guedes J, Pegon P, Pinto AV (1994) A fibre/Timoshenko beam element in Castem 2000. Joint Research Centre
Special publication No. I.94.31, European Commission, Ispra, Italy
Hoshikuma J, Unjoh S, Nagaya K (2001) Size effect on ductile behaviour of reinforced concrete columns
under cyclic loading. In: Proceeding of the 17th USA—Japan Bridge Engineering Workshop, Tsukuba,
Japan
International Federation for Structural Concrete (fib) (2006) Seismic bridge design and retrofit—structural
solutions: state-of-the-art report. Bulletin 39. Lausanne, Switzerland
Jacobsen LS (1930) Steady forced vibrations as influenced by damping. ASME Trans 52:169–181
Loeding S, Kowalsky MJ, Priestley MJN (1988) Direct displacement-based design of reinforced concrete
building frames. University of California, San Diego
Mander JB, Priestley MJN, Park R (1988) Theoretical stress-strain model for confined concrete. J Struct Eng
ASCE 114(8):1804–1826
Menegotto M, Pinto PE (1973) Method of analysis for cyclically loaded reinforced concrete plane frames
including changes in geometry and non-elastic behaviour of elements under combined normal force and
bending. In: Proceedings of the IABSE Symposium on resistance and ultimate deformability of structures
acted on by well-defined repeated loads, Lisbon, Portugal
Mercer C, Martin J (1987) A beam element for cyclically loaded reinforced concrete structures. Technical
Report No. 98, FRD/UCT Centre for Research in Computational and Applied Mechanics, University of
Cape Town, South Africa
Monti G, Nuti C (1992) Nonlinear cyclic behaviour of reinforcing bars including buckling. J Struct Eng, ASCE
118(112):3268–3284
Park R, Priestley MJN, Gill WD (1982) Ductility of square-confined concrete columns. J Struct Eng Div
108(ST4):929–950
Paulotto C, Ayala G., Taucer F, Pinto A (2007) Simplified models/procedures for estimation of secant-to-
yielding stiffness, equivalent damping, ultimate deformations and shear capacity of bridge piers on the
basis of numerical analysis. JRC Scientific and Technical Report EUR22885EN, Joint Research Centre,
European Commission, Ispra, Italy

123
1420 Bull Earthquake Eng (2010) 8:1397–1420

Pinto AV, Verzelletti G, Negro P, Guedes J. Cyclic testing of a squat bridge-pier (1995) JRC Scientific and
Technical Report EUR16247EN, Joint Research Centre, European Commission, Ispra, Italy
Pinto AV, Verzelletti G, Pegon P, Magonette G, Negro, P, Guedes J (1996) Pseudo-dynamic testing of large-
scale R/C bridges. JRC Scientific and Technical Report EUR16378EN, Joint Research Centre, European
Commission, Ispra, Italy
Priestley JN, Kowalsky MJ (1998) Aspects of drift and ductility capacity of rectangular cantilever structural
walls. Bulletin New Zealand National Society for Earthquake Engineering 31(6):73–85
Priestley MJN, Seible F, Calvi GM (1996) Seismic design and retrofit of bridges. Wiley, New York
Stone WC, Cheok GS (1989) Inelastic behaviour of full-scale bridge columns subjected to cyclic loading.
NIST Building Science Series 166, National Institute of Standards and Technology, US Department of
Commerce, Washington

123

You might also like