You are on page 1of 14

An Efficient Methodology to Estimate Probabilistic

Seismic Damage Curves


Yeudy F. Vargas-Alzate 1; Luis G. Pujades 2; Alex H. Barbat 3; and Jorge E. Hurtado 4

Abstract: The incremental dynamic analysis (IDA) is a powerful methodology that can be easily extended for calculating probabilistic
seismic damage curves. These curves are metadata to assess the seismic risk of structures. Although this methodology requires a relevant
computational effort, it should be the reference to correctly estimate the seismic risk of structures. Nevertheless, it would be of high practical
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

interest to have a simpler methodology, based for instance on the pushover analysis (PA), to obtain similar results to those based on IDA.
In this article, PA is used to obtain probabilistic seismic damage curves from the stiffness degradation and the energy of the nonlinear part of
the capacity curve. A fully probabilistic methodology is tackled by means of Monte Carlo simulations with the purpose of establishing that
the results based on the simplified proposed approach are compatible with those obtained with the IDA. Comparisons between the results of
both approaches are included for a low- to midrise reinforced concrete building. The proposed methodology significantly reduces the com-
putational effort when calculating probabilistic seismic damage curves. DOI: 10.1061/(ASCE)ST.1943-541X.0002290. © 2019 American
Society of Civil Engineers.

Introduction and (3) the uncertainties related to the structural properties and
to the seismic hazard. These aspects allow to obtain probabilistic
In the last decades, several approaches have been developed for damage curves (PDC), which are metadata for estimating seismic
estimating the seismic risk of structures. One of these is based risk scenarios at an urban level. Notice that to correctly estimate
on the vulnerability index method. In this method, the seismic haz- seismic risk scenarios, the uncertainties involved in the seismic
ard is considered by means of macroseismic intensities (Egozcue hazard action and in the mechanical and geometrical properties
et al. 1991), and the structural behavior by means of vulnerability of the structures should be considered. Nonetheless, the computa-
indices (Lantada et al. 2010; Barbat et al. 2011; Lagomarsino and tional effort involved to carry out IDA can be very large, and the use
Giovinazzi 2006). Another widely invoked method is based on of simplified methods is an urgent need of practical interest.
the capacity spectrum method (CSM) developed by Freeman In this paper a simplified method for calculating PDC, with
et al. (1975), Freeman (1998), and further developed by Fajfar and an efficient approach based on the stiffness degradation and on
Gaspersic (1996), Chopra and Goel (1999), and Fajfar (1999), the energy of the nonlinear part of the capacity curve, is proposed.
among others. In this method, the seismic action is considered by The PDC obtained with this simplified approach fit with sufficient
means of the elastic response spectrum, while the building capacity accuracy the PDC calculated via IDA method. The effectiveness of
is represented by means of a capacity curve or a capacity spectrum. the simplified presented approach is tested for a low- to midrise
The capacity curve is calculated with an incremental nonlinear reinforced concrete framed building. The Monte Carlo method
static analysis, commonly known as pushover analysis (PA); it
is used to perform the probabilistic calculations presented in this
is also assumed that the structural response is dominated by the
paper.
first mode of vibration. However, note that the capacity spectrum
method has been further modified in order to include the effects of
higher modes of vibration (Chopra and Goel 2002; Chopra et al.
2004; Kreslin and Fajfar 2012). More recently, the incremental Incremental Dynamic Analysis
dynamic analysis method (IDA) has become a powerful tool for A nonlinear dynamic analysis (NLDA), allows to simulate the time
assessing the seismic behavior of buildings, which allows consid- history response of a structure subjected to an earthquake. This
eration of the following aspects: (1) the nonlinear dynamic re-
analysis also allows to calculate the maxima of structural response
sponse of the structure; (2) the increments of the seismic intensity;
variables like the displacement at the roof, the global damage index,
1 among others. A method that uses NLDA as structural solver is
Researcher, Dept. of Civil and Environmental Engineering, Univer-
sidad Politécnica de Cataluña, Jordi Girona 1-3, Barcelona 08034, Spain
the IDA (Vamvatsikos and Cornell 2002). This method is generally
(corresponding author). Email: yeudy.felipe.vargas@upc.edu performed with several earthquakes records increasingly scaled.
2
Full Professor, Dept. of Civil and Environmental Engineering, Univer- The IDA method may be extended to include uncertainties in the
sidad Politécnica de Cataluña, Jordi Girona 1-3, Barcelona 08034, Spain. structural properties (Vamvatsikos and Fragiadakis 2010; Vargas
3
Full Professor, Dept. of Civil and Environmental Engineering, Univer- et al. 2012), ensuing a very good approach to adequately estimate
sidad Politécnica de Cataluña, Jordi Girona 1-3, Barcelona 08034, Spain. PDC. Groups of PDC are generally used to simulate seismic risk
4
Full Professor, Dept. of Civil Engineering, Universidad Nacional de scenarios at urban areas. However, if uncertainties are included
Colombia, Manizales 170002, Colombia.
when employing IDA, it generally turns into a computationally
Note. This manuscript was submitted on October 13, 2017; approved
on September 18, 2018; published online on January 28, 2019. Discussion costly method (Vargas et al. 2015). In spite of this, this section
period open until June 28, 2019; separate discussions must be submitted is devoted to explaining how IDA can be applied for obtaining
for individual papers. This paper is part of the Journal of Structural PDC. A low- to midrise reinforced building is considered as a
Engineering, © ASCE, ISSN 0733-9445. testbed.

© ASCE 04019010-1 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Fig. 1. (a) Building studied in this paper; and (b) 2D model.
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

Table 1. Characteristics of the elements of the studied building shown in Table 2. Characteristics of the input random variables
Fig. 1
Variable μ σ COV
Columns Beams 4 3
fc (kPa) 2.1 × 10 3.15 × 10 0.15
Story b (m) h (m) ρ b (m) h (m) ρ Es (kPa) 2 × 108 2 × 107 0.1
1 0.5 0.5 0.03 0.45 0.6 0.0066 Note: μ, σ, and COV represent mean, standard deviation, and coefficient of
2 0.5 0.5 0.02 0.45 0.6 0.0066 variation of the random variables, respectively.
3 0.45 0.45 0.015 0.45 0.6 0.0066
4 0.4 0.4 0.015 0.45 0.6 0.0066
Note: b, h, and ρ denote base, height, and steel percentage of the cross (COV) of these random variables. Truncated normal distributions
sections of the structural elements, respectively. are considered, being the limits the mean value minus and plus 2.5
standard deviations. Thus, negative samples are not allowed for var-
iables in Table 2. The coefficient of variation of the concrete com-
Description of the Studied Building pressive strength, fc, may vary from building to building, in the
range of 0.07–0.2; the average value of this COV is about 0.15
A reinforced concrete framed building has been designed and used
(Melchers 1999). However, this value depends on the quality control
as a testbed in order to perform the structural calculations proposed
carried out during construction. For instance, a highly controlled
in this paper. A sketch of this building is shown in Fig. 1(a). This
concrete tends to exhibit a COV value close to 0.1 (Melchers 1999).
figure also displays the geometrical dimensions of the building.
For a midcontrolled concrete, the expected COV is about 0.15. In
Notice that, due to its symmetry, the building can be modeled as a
this study, the COV for the concrete compressive strength has been
two-dimensional (2D) structure by using a single frame [Fig. 1(b)].
assumed to be 0.15. The coefficient of variation of the elastic modu-
The primary characteristics of the beams and columns of the
lus of the steel, Es, has been set to 0.1 (Melchers 1999).
framed structure are given in Table 1. The inelastic behavior of
Other likely uncertainties such as those related to cracking and
beams and columns elements follow the concept of the Gilberson
crushing of concrete, strain hardening, and ultimate strength of
one-component model, which allows hinges at one or both ends of
steel, and effects such as slab participation or uncertainties related
the elastic central length of the member. In this model, the length of
to the model options, can be also included in the probabilistic struc-
plastic hinges was assumed 5% of the length, L, of the element; that tural analysis. Note that if new random variables were considered
is 0.05 L. To consider the hysteretic cycle for flexure-failing they would affect both dynamic and static results. In this paper, the
mechanism in beams and columns, the modified Takeda hysteresis random variables that mostly affect the uncertainties of the re-
law (Otani 1974) has been considered. Yielding surfaces are de- sponse in an explicit way were considered. Nonetheless, notice that
fined by the bending moment-axial load interaction diagram for the uncertainties of several structural variables are taken into ac-
columns and bending moment-curvature for beams. The applied count indirectly as they are calculated as functions of the explicitly
loads follow the recommendations given by Eurocode 2 (CEN considered random variables. For instance, the elastic modulus
2004a) for reinforced concrete structures. Notice that nonlinearities p of
the concrete, Ec, is related to fc by the equation Ec ¼ 4,500 fc,
due to shear forces are not allowed because it has been assumed that and for the calculation of the yield strength, fy, the expression
the structural elements are well confined. fy ¼ 0.0021Es is used.

Mechanical Properties of the Materials as Random Seismic Hazard to Perform IDA


Variables
One of the most important sources of uncertainty when estimat-
Generally, the characteristic mechanical properties of concrete and ing the seismic risk of structures is the random variability of the
steel are the values used in the design of reinforced concrete build- ground-motion. Its influence has been studied by Bommer and
ings. Design standards require characteristic strength values for the Crowley (2006). The forecasting of the ground-motion parameters
materials obtained during the quality control process, from com- has been studied by Abrahamson et al. (1991), Bommer et al.
pression and tension tests in concrete and steel samples, respec- (2007), and Arroyo and Ordaz (2011). According to the probabi-
tively. In order to meet the probabilistic approach of this paper, the listic simulation approach of this study, it is necessary to model the
concrete compressive strength, fc, and the steel elastic modulus, seismic hazard as a random variable. To do that, 10 earthquakes
Es, will be modeled as random variables. Table 2 presents the as- have been selected from the European database (Ambraseys et al.
sumed mean, μ, standard deviation, σ, and coefficient of variation 2004) by using a criterion for controlling the dispersion of the data.

© ASCE 04019010-2 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


the most similar to the target spectrum is ranked as the first one,
and so on. Note that this selection procedure avoids the sensitivity
of the method to single records, as the sorting procedure itself dis-
cards records leading to likely significant differences in the dynamic
analyses as well as in the computations of the performance points. It
was also found that the number of records that minimize the error
function as defined in Vargas et al. (2013a) takes values between 10
and 30. A total of 10 records have been used to take into account the
randomness of the seismic actions in this study.
However, an anonymous reviewer noted that the way the seis-
mic actions have been selected could underestimate the record-to-
record variability, thus enhancing the variability in the response due
to the randomness of the mechanical properties. Eads et al. (2013)
quantified the uncertainty in the collapse fragility curves and mean
annual frequency of collapse as a function of the number of ground
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

motions used in calculations, showing that using a small number of


ground motions can lead to unreliable estimates of structure col-
lapse risk. The importance of this issue in seismic expected damage
and seismic risk assessments is recognized and should be faced
Fig. 2. Response spectra calculated from the procedure based on the carefully in future research. The record selection method used in
mean spectrum. this paper needs to be checked against other ground motion selec-
tion methods, for instance the one proposed by Lin et al. (2013a)

μDepth ¼ 11 km; σDepth ¼ 3.8 km; μED ¼ 22.9 km;


Use is made of a method proposed by Vargas et al. (2013a) for
σED ¼ 7.65 km; μMs ¼ 6.13; σMs ¼ 0.55
obtaining a set of earthquake records from a database. Such records
fit a target response spectrum under a mean square metric. The re- Note that, in addition of finding earthquakes records compatible
sponse spectral function selected in this study corresponds to Type with a target spectrum, the method proposed by Vargas et al.
1 of the Eurocode 8 (EC8) (CEN 2004b), which has a surface-wave (2013a) also has the ability to find and to select records correspond-
magnitude greater than 5.5. Several methods to establish the opti- ing to similar earthquake features, including depth and epicentral
mal number of accelerograms required to perform the inelastic distance, local geology, and size of the earthquakes, according to
dynamic analyses have been studied by Hancock et al. (2008), the great (M S > 5.5) and moderate (M S <¼ 5.5) earthquakes, as
and they conclude that the exact number depends on the damage considered in Eurocode 8. At the end of Table 3, statistics of these
measures that are considered and also on their predictability. Most variables are shown. Mainly shallow great near earthquakes, re-
of these methods are based on the magnitude and on the spectral corded at rock or stiff soils, were automatically selected. Note
shape. Fig. 2 depicts the spectra of the selected earthquakes, their also that dispersions around the central values are relatively small.
average spectrum, and the target spectrum. Fig. 2 shows that the Noticeably, this method for selecting the accelerograms also guar-
procedure used for selecting earthquake records allows a good fit. antees some record-to-record variability; in fact, for the entire range
The list of these earthquakes and of their primary characteristics are of periods, the coefficients of variation of the spectral acceleration
given in Table 3. values are in the range between 0.16 and 0.2, and as no matching
The selection of the earthquake records according to the method techniques are used, the frequency content, amplitudes, and phases
proposed by Vargas et al. (2013a) is based on the mean squared are retained. Lin et al. (2013a, b) analyzed the crucial problem of
error calculated between the mean of a group of response spectra ground motion selection based on a target spectrum, concluding
previously selected from a sorted database criterion and the target that the use of uniform hazard spectrum, including variable uncer-
spectrum function. The criterion for sorting the database is based on tainties at different periods, results in overestimations of structural
the mean square errors between actual spectra and a target spec- response hazard, whereas the conditional mean spectrum results
trum. The earthquake records are sorted according to such errors in underestimation. Thus, the uncertainties due to seismic actions
in ascending order; thus, the record whose response spectrum is could be underestimated in this study.

Table 3. Main characteristics of the selected earthquakes


Epicenter (degrees)
Depth Magnitude Local Epicentral
Station name Date N E (km) (Ms) geology distance (km)
Arquata del Tronto September 19, 1979 42.76 13.02 4 5.84 Rock 22
San Rocco September 15, 1976 46.32 13.16 12 5.98 Stiff soil 17
Kotor NasRakit May 24, 1979 42.23 18.76 5 6.34 Rock 21
Auleta November 23, 1980 40.78 15.33 16 6.87 Rock 25
Ponte Corvo May 7, 1984 41.73 13.90 8 5.79 Rock 31
Matelica September 26, 1997 43.03 12.86 6 5.9 Rock 20
Tricarico May 5, 1990 40.65 15.92 12 5.6 Rock 20
Izmit-M-Istasyonu September 13, 1999 40.70 30.02 13 5.9 Stiff soil 13
Bolu-Bayindirlik November 12, 1999 40.76 31.14 14 7.3 Stiff soil 39
Athens-Papagos September 7, 1999 38.13 23.54 9 5.6 Rock 26

© ASCE 04019010-3 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Probabilistic Calculation Based on IDA A brief description of the dynamic damage index employed in this
study is given in the following section.
To take into account the uncertainties related to the variables de-
scribed in Table 2, the Monte Carlo method is used. Nonetheless,
an important issue related to the spatial variability should be treated Park and Ang Damage Index
before starting the calculation. It is well known that the spatial vari-
ability of the mechanical features of the structural elements may Several damage indices for reinforced concrete elements from a
influence greatly the results (Franchin et al. 2010). For instance, postprocess of the nonlinear dynamic response have been proposed
in order to consider the likely correlation that may exhibit the in the literature (Williams and Sexsmith 1995). One of the most
strength of concrete for the elements of a specific story, it is nec- well established is the damage index of Park and Ang (1985),
essary to assume a correlation hypothesis between the random gen- which is calculated as the sum of the maximum ductility divided
erated samples. Thus, it is supposed that the correlation between the by the ultimate ductility and a term related to the dissipated energy.
strength of columns and beams of the same story decreases with In the case of the overall building, ductility is defined in terms of
distance. To do this for the 2D frame depicted in Fig. 1, the column displacement, but for several structural members ductility can be
of the first story numbered as k ¼ 1 will be the first from left to defined in terms of other parameters, such as plastic rotation. The
right, the column numbered as k ¼ 2 of the first story will be the corresponding equation is
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

second one going from left to right, and so on. Thus, the correlation μm ðχÞ β h Eh ðχÞ
matrix, ρi;j , for the concrete strength samples of one story is made DI E ðχÞ ¼ þ ð2Þ
in the following way: μu F y μu δ y
8 where DI E = damage index at element level; and χ is an indepen-
<i ¼ j ρi;j ¼ 1
dent variable, depending of the severity of the seismic action or,
ρi;j ¼ ð1Þ
: i ¼ j  k ρi;j ¼ 1 − k equivalently, of some parameter linked to the structural response.
r In this case, the PGA is used as the variable χ. Other variables, such
as roof displacement and/or spectral displacement, can be also used
where r is a coefficient associated to the level of correlation be-
as χ variable. In this case, μm and μu are the maximum and ultimate
tween adjacent elements. Further details of the hypothesis assumed
ductilities, respectively; and β h is a nonnegative parameter that rep-
to calculate the correlation matrix presented in Eq. (1) can be found
resents the effect of cyclic loads on the structural damage. Taking
in Vargas et al. (2017). Based on this correlation matrix, a group
into account the calibration of the Park and Ang damage index for
of correlated random samples can be generated for the concrete
reinforced concrete buildings, the β h value has been set to 0.05; Eh
strength of columns of each story, while for each column of the
is the hysteretic energy dissipated; Fy is the yield force; and δ y is
same story, one independent random sample of the elastic modulus
the yield displacement. Then, the global damage index of the struc-
of the steel, Es, is assigned. The same criterion is used to generate
ture, DI NLDA , is a weighted mean of the damaged elements, in
random samples of the mechanical characteristics of the beams for
which the weights are the ratio of the hysteretic energy dissipated
each story. Note that groups of samples corresponding to different
by each element to the total hysteretic energy dissipated by the struc-
stories are independent. These considerations are based on the fact
ture. The following equation describes this global damage index:
that usually structural elements of the same story are made of the
same concrete but the properties of the reinforcement steel, from X
DI NLDA ðχÞ ¼ λi DI E ðχÞ ð3Þ
rebar to rebar, can be supposed as independent. Recall that the aim i
of this paper is related to the assessment of the expected seismic
damage of individual buildings; nevertheless, this purpose could be where DI NLDA ðχÞ = dynamic-analysis-based global damage index
extended to classes of existing structures in urban areas taking into of the structure; and λi = ratio of the dissipated hysteretic energy
account the building-to-building variation of the structural charac- of an element E to the dissipated hysteretic energy of the entire
teristics within a structural class. Further discussion on this issue structure. Recall that the Park and Ang index provides a measure of
can be found in Crowley et al. (2005). the global damage of the structure by considering two sources of
Consequently, according to the spatial variability hypothesis damage: the ductility reached for each structural element, and the
presented, random samples of the mechanical properties of the dissipated energy due to the hysteresis. The maximum ductility is
materials are generated, and then NLDAs are performed for the related to the stiffness degradation, which is generally the primary
earthquakes records linearly scaled. Because of the probabilistic contributor to damage in reinforced concrete elements, but the con-
approach, the fundamental period varies from building to building. tribution to damage of dissipated energy is also significant. More-
Thus, the intensity measure (IM), selected for scaling the records over, dissipated energy is probably the best way of taking into
has been the peak ground acceleration (PGA). However, as also account the damaging effects of the earthquake duration.
pointed out by one anonymous reviewer, other options could be fea- Based on the results obtained by means of IDA, PDC relating
sible, or even preferable. For instance, the spectral acceleration re- the PGA to the damage index can be obtained. Fig. 3(a) displays
lated to the mean period from all models of the Monte Carlo, or the such PDC. According to the original calibration of the damage in-
period of the model having the mean (or median) parameter values, dex made by Park and Ang (1985), a damage index value equal or
could be good options for scaling the seismic records. However, to higher than 1 corresponds to the collapse of the structure. For this
compare and discuss the influence of the different scaling options in reason, when the damage index of Park and Ang is higher than 1, its
probabilistic approaches requires additional work that falls beyond value is set to 1. This makes sense because for higher PGAs, the
the primary scope of this study. Therefore, PGA values from 0.05g standard deviation of the damage index decreases, which indicates
to 1g, at intervals of 0.05 g, have been considered. The Latin hyper- that for higher PGAs the collapse uncertainty reduces.
cube method is used to randomly combine earthquakes records and Notice that results presented in Fig. 3(a) can be used to the
simulated buildings for performing NLDAs. A total of 1,000 estimation of failure probability functions or even fragility func-
NLDAs are performed for each PGA. In each analysis, the seismic tions in terms of earthquake intensities. For instance, in Fig. 3(b)
global damage index of the building is obtained as a weighted mean the fragility functions are presented by assuming that the damage
of the damage indices initially estimated at structural element level. states thresholds for slight, moderate, extensive, and complete are

© ASCE 04019010-4 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. (a) Park and Ang damage indices represented as a function of the PGA for the building shown in Fig. 1; and (b) fragility functions.

0.1, 0.4, 0.65, and 1, respectively. This assumption is based on the


original calibration of the damage index in Park et al. (1985).
To summarize the simulation scheme for calculating PDC based
on IDA, a flow chart has been produced, and is shown in Fig. 4.
As mentioned previously, seismic intensity is defined by PGA
values in the range between 0.05g and 1g. Increments of 0.05g are
considered, so 20 PGA increasing values are used in the IDA. The
following steps summarize the simulation scheme:
1. Generate a random structural model (Counter i, i ¼ 1∶100).
2. Take the earthquake record of Table 3 corresponding to the
iteration number (Counter j, j ¼ 1∶10).
3. Increase the seismic intensity and scale the record to this PGA
value (Counter k, k ¼ 1∶20).
4. Compute a NLDA and calculate the Park and Ang damage
index, DI NLDA . Store the DI NLDA calculated in the matrix
MRDYN ð10 × ði − 1Þ þ j; kÞ ¼ DI NLDA ; check for intensity
PGA. If the PGA is lower than the maximum PGA considered,
go to Step 3; otherwise check for record number. If the record
number is lower than the maximum number of records con-
sidered, then go to Step 2; otherwise check for the DI NLDA
statistical distribution. Then the convergence of the two first
statistical moments of the distribution of DI NLDA is checked.
If this convergence is unsatisfactory, then go to Step 1; other-
wise end.
Note that in the present study a number of 100 structural models
together with 10 seismic actions warrants this convergence; thus,
this last check is avoided.

Damage Index Based on Pushover Analysis

As mentioned previously, when uncertainties of the material proper-


ties and of the seismic action are considered, IDA is an expensive
computational way for obtaining PDC. Vargas et al. (2013a, b) have
shown a simplified procedure to obtain PDC based on the pushover
analysis and fragility curves. However, this procedure does not al-
low separation of the damage into two components, like the damage
index of Park and Ang approach. It is of high practical interest to
develop a computationally less expensive procedure, based for in-
stance on nonlinear static structural analysis, which allows obtaining
PDC, similar to those calculated with the IDA method presented
previously. To do this, the incremental static analysis (i.e., pushover
Fig. 4. Flow chart of the simulation scheme for calculating probabil-
analysis) combined appropriately with an innovative procedure for
istic damage curves based on NLDA.
estimating damage curves (Pujades et al. 2015) provides a powerful

© ASCE 04019010-5 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


and simplified tool. To calibrate the pushover-based damage index, consuming, as it requires evaluating the next pushing pattern in
the global damage index of Park and Ang was selected because, as it each iteration of the pushover analysis. A detailed description of
will be shown subsequently, both damage indices represent overall this procedure can be found in the RUAUMOKO’s manuals (Carr
damage in the building. However, the global index is a general in- 2000). This structural program has been used in this study for cal-
dicator of the damage and is the result of a weighted average of the culating static and dynamic nonlinear structural analysis.
evolution of damage at a component level. According to the values exposed in Tables 1, 100 groups of
Gaussian input data samples are generated according to the cor-
relation hypotheses exposed previously (section “Probabilistic cal-
Probabilistic Pushover Analysis
culation based on IDA”); then, the adaptive pushover analyses are
The PA consists in applying a horizontal load to the structure performed. Fig. 5 depicts the 100 capacity curves obtained, which
according to a certain pattern of forces. This load value is increased are characterized by random variables such as the elastic stiffness,
until the structural collapse is reached. Afterward, a curve relat- the maximum displacement, and the maximum base shear. Notice
ing the displacement at the roof to the base shear of the building that the 100 capacity curves shown in Fig. 5 can be easily trans-
is obtained. This curve is commonly known as capacity curve. formed into 100 capacity spectra as per HAZUS (FEMA 1999) and
Several researchers have shown the primary advantages and disad- ATC-40 (ATC 1996). The right and top axes in Fig. 5 show this
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

vantages when PA methods are employed to calculate structural transformation.


reliability estimates (Fragiadakis and Vamvatsikos 2010; Celarec
and Dolšek 2013; and Brozovič and Dolšek 2014). Notice that
Calculation of the Performance Points by Considering
if the PA will be used in a probabilistic environment, several issues
Uncertainties
should be considered. In this way, Satyarno (2000) proposed the
adaptive pushover analysis. An important advantage of the adaptive From a capacity spectrum of a structure and a response spectrum,
pushover is that the shape of the lateral load in each incremental there are several simplified methods to estimate the structural
load step is recalculated based on the first mode of vibration of seismic response, namely HAZUS (FEMA 1999), ATC-40 (ATC
the degraded structure. Besides, a Gordian knot in pushover analy- 1996), and FEMA 440 (FEMA 2005). These simplified methods
sis is to take a decision on the ultimate capacity point. Although generally look to obtain the performance point. This latter is a mea-
several criterions exist, based on the decay of the maximum capac- sure of the expected demand of a structure. Thus, in order to cal-
ity, the adaptive pushover automatically ends pushing based on the culate the performance point of the analyzed building, due to the
stiffness degradation. Namely, when the eigenfrequency of the same seismic hazard used in the IDA calculations, the response
degraded structure is less than 10−6 times the initial elastic eigen- spectra of the earthquake records presented in Table 3 should be
frequency, the method automatically stops incrementing the lateral calculated. These response spectra, together with the capacity spec-
load. In fact, it is assumed that a failure mechanism has occurred tra depicted in Fig. 5, are displayed in Fig. 6. Obviously, the re-
and that further increments would provide unrealistic values. sponse spectra should be scaled to different intensity levels in
These advantages are more important in probabilistic pushover the same way as in the IDA method. So, to facilitate the comparison
analysis because of the automatic solution of these two important of the results, the intensity levels will be the same ones that were
concerns. The primary limitation of this procedure is that it is time used in the IDA calculations; that is, PGAs ranging from 0.05g to

Fig. 5. Capacity curves and capacity spectra of the building studied by considering the mechanical properties of the materials as random variables.

© ASCE 04019010-6 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Performance point for a given demand and capacity spectra. The
method applied is exposed in Chapter 6 of FEMA 440 (FEMA 2005).

Fig. 7. Relation between PGA to spectral displacement.

1g at intervals of 0.05g. Fig. 6 also shows an example of calculation


Fig. 8. Flow chart of the simulation scheme for calculating capacity
of the performance points, for a PGA of 1 g. In this figure the in-
curves and the performance point PP by considering the response
elastic spectra, reduced due to the ductility of the building, are also
spectra for the earthquakes records shown in Table 3.
shown. Note that the spectral displacement of the performance
point is also referred to as target displacement in the international
literature. The methodology applied for obtaining the performance
point is further explained in FEMA 440 (FEMA 2005).
After computing all the performance points (20,000 in this case), 3. Increase the seismic intensity and scale the spectrum to this
1,000 curves relating PGAs and the spectral displacements of the per- PGA value (Counter k, k ¼ 1∶20).
formance points are obtained. Fig. 7 displays these curves, which will 4. Compute the performance point and store the spectral dis-
be useful to easily combine the uncertainties related to the mechanical placement of the performance point, SdPP , in the matrix
properties of the materials and those related to the seismic hazard. MRSd-PGA ð10 × ði − 1Þ þ j; kÞ ¼ SdPP . Check for intensity PGA.
To summarize the simulation scheme for calculating the curves If the PGA is lower than the maximum PGA considered, go to
of Fig. 7, a flow chart is depicted in Fig. 8. The primary steps of the Step 3; otherwise check for record number. If the record number
proposed approach are as follows: is lower than the maximum number of records considered, then
1. A capacity curve is computed for a randomly generated struc- go to Step 2; otherwise check for the SdPP statistical distribution.
tural model (Counter i, i ¼ 1∶100). Check the convergence of the two first statistical moments of the
2. Take the response spectrum of the record of Table 3 correspond- distribution of SdPP . If this convergence is unsatisfactory, then go
ing to the iteration number (Counter j, j ¼ 1∶10). to Step 1; otherwise end.

© ASCE 04019010-7 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Note that a number of 100 structural models together with 10 ke − ks ðSdÞ
DI SS ðSdÞ ¼ ð6Þ
seismic actions warranties this convergence; thus, this last check is maxðke − ks ðSdÞÞ
avoided.

where DI TS ðSdÞ = damage index based on the tangent stiffness;


Calculation of the New Simplified Damage Index ke = elastic stiffness; kt ðSdÞ = tangent stiffness; DI SS ðSdÞ =
As stated previously, the Park and Ang damage index is the sum of damage index based on the secant stiffness; and ks ðSdÞ = secant
the contributions to damage of the maximum ductility and of the stiffness. Observe that both DI TS ðSdÞ and DI SS ðSdÞ have been nor-
energy dissipated by the structural elements. Pujades et al. (2015) malized by maxðke − kt ðSdÞÞ and maxðke − ks ðSdÞÞ, respectively,
proposed an analogue equation for estimating a seismic damage in such a way that their highest value is 1.
index based on the PA The next step consists in selecting which of the calculated stiff-
ness, tangent or secant, is a better candidate to represent the first
DI PA ðSdÞ ¼ αDI μ ðSdÞ þ βDI DE ðSdÞ ð4Þ term of Eq. (4). If the first term of the Park and Ang equation is
analyzed in detail, one can see that the maximum ductility reached
where DI PA ðSdÞ = pushover analysis–based global damage index during the time history response is normalized by the ultimate duc-
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

of the structure; α is a coefficient that weights the contribution to tility. This fact indicates that the rate of evolution of the ductility
the global damage of the first term of DI PA ðSdÞ related to the index is closer related to the secant stiffness than to the tangent stiff-
ductility; DI μ Sd = damage term related to the ductility; β is a co- ness. Besides, the curves of Fig. 5 do not show strong and sudden
efficient that weights the contribution to the global damage of the losses of stiffness. The later behavior can be likely in rigid struc-
second term of DI PA ðSdÞ; and DI DE ðSdÞ = damage term related to tures, for which it can be expected that a combination of the tangent
the dissipated energy. and the secant stiffness should be considered. For instance, if the
To obtain two functions from the capacity spectrum, to represent infill walls were modeled nonlinearly it would certainly lead to sud-
adequately the terms of Eq. (4), it is necessary to analyze the tan- den losses of stiffness and strength, and a linear combination be-
gent and secant stiffness functions as well as the dissipated energy tween the secant and the tangent stiffness will probably be a better
curve; as it will be seen further on, these curves are obtained easily option for the first term of Eq. (4). Further research on this issue will
from the capacity spectrum. As shown previously, the first term of allow generalizing Eq. (4) for new structural typologies. However,
the equation of Park and Ang [Eq. (2)] is related to the maximum as mentioned previously, for the building types considered in this
ductility reached by a structural element and it can be related to the paper, it is expected that the major contribution to damage be due
stiffness degradation of the building. In the case of the capacity to secant stiffness degradation. Thus, the damage index based on
spectrum method, the first derivative of the capacity spectrum is the secant stiffness has been selected as the first term of Eq. (4),
directly related to the stiffness degradation. However, two kinds of i.e., DI μ ðSdÞ ¼ DI SS ðSdÞ.
stiffness degradation can be calculated: the first one based on the The second term of the proposed damage index in Eq. (4) stands
tangent stiffness, and the second one on the secant stiffness. for the dissipated energy as a function of the spectral displacement.
Fig. 9(a) shows the tangent and secant stiffness of the capacity This term is obtained as the cumulative integral of the nonlinear
spectra depicted in Fig. 5. part of the capacity curve as defined in Pujades et al. (2015). This
These stiffness functions can be transformed into a damage cumulative integral also can be easily obtained by integrating the
index based on the tangent stiffness or into a damage index based function defined as the subtraction of the capacity curve, SaðSdÞ,
on the secant stiffness by means of the following equations [see from a projected line of the elastic part of the capacity spectrum.
Fig. 9(b)]: Then, this energy-based index is normalized so that its value at the
ultimate capacity point is 1. Thereby, this index holds for the rel-
ke − kt ðSdÞ
DI TS ðSdÞ ¼ ð5Þ ative energy dissipated as referred to the energy loss at the ultimate
maxðke − kt ðSdÞÞ or collapse point. The following equation is used to compute this

Fig. 9. Tangent stiffness and secant stiffness of the capacity spectrum.

© ASCE 04019010-8 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Graph depicting the functions and areas involved in the
definition of the energy term in Eq. (4) as described and formulated
in Eq. (7).
Fig. 12. New damage index for α ¼ 0.5.

Fig. 11. Stiffness and energy functions for calculating the new damage
index.
Fig. 13. Comparison between damage indices by using IDA and PA.

energy term and Fig. 10 illustrates the meaning of the involved


functions and areas However, these curves do not take into account the seismic
R ξ¼Sd action yet. Thus, to take account of it, a change of variable is per-
ξ¼0 ðke  ξ − SaðξÞÞdξ formed by transforming the spectral displacement into a parameter
DI DE ðSdÞ ¼ R ξ¼Sd ð7Þ
max½ ξ¼0 ðke  ξ − SaðξÞÞdξ explicitly linked to the seismic hazard, for instance PGA. It is im-
portant to recall that 1,000 curves, connecting PGA and spectral
Fig. 11 depicts both terms of Eq. (4), DI μ ðSdÞ and DI DE ðSdÞ. displacement for the building and the seismic actions, were shown
According to Eq. (4), the coefficient α and β are weights of the in Fig. 7. These functions allow making the appropriate change of
damage terms related to the ductility and energy functions, respec- variable, as each of these curves were obtained by calculating the
tively. However, it is important to recall that, in this paper, the val- performance points in an incremental way. Then, the new damage
ues of the Park and Ang damage index, when larger than 1 were set index can be expressed as a function of the peak ground accelera-
to 1. Pujades et al. (2015) proposed the following assumption to tion. It is important to note that these performance point computa-
avoid that the values of the new damage index be larger than 1: tions have been made by using the same PGA values that were used
α¼1−β ð8Þ in the nonlinear dynamic analysis.
In Fig. 13, the PDC calculated by using IDA are compared with
Therefore, Eq. (8) allows rewriting Eq. (4) as follows: the probabilistic damage curves calculated by using PA for an
DI PA ðSdÞ ¼ αDI μ ðSdÞ þ ð1 − αÞDI DE ðSdÞ ð9Þ α ¼ 0.7. Even if this comparison seems to exhibit a relatively good
fit among the Park and Ang damage index and the proposed one,
If Eq. (9) is applied for α ¼ 0.5, the curves shown in Fig. 12 are important differences appear when likening the percentiles of each
obtained. group of PDC. In order to obtain a coefficient α, which provides the

© ASCE 04019010-9 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


best fit between the curves obtained via IDA and PA, the mean
square error (MSE) for percentiles 5, 50, and 95 are calculated
for a suite of α values. Fig. 14 shows the evolution of the MSE
with α, for each percentile. α values go from 0 to 1, with incre-
ments of 0.01.
Fig. 15 shows the comparison between the curves obtained by
using IDA and PA approaches for percentiles 5, 50, and 95. The
best fit is obtained by using α ¼ 0.65, α ¼ 0.78, and α ¼ 0.67 for
percentiles 5, 50, and 95, respectively. It is important to observe the
good adjustment exhibited by the curves calculated with both meth-
ods. This fact demonstrates the validity of the simplified proposed
approach based on PA.
However, it is also true that α is an important parameter and that
its value is critical to calculate the new damage index. Thus, a key
issue concerning to this new damage index is that parameter α is
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

calibrated on the basis of the Park and Ang index. Indeed, this
parameter is selected using a criterion based on the error with
respect to the reference solution. So the advantage of the new pro-
Fig. 14. Mean square error for percentiles 5, 50, and 95 as a func-
cedure seems to fade away. Moreover, one would expect that
tion of α.
this error should not be always available, particularly for involved

Fig. 15. Comparison between percentile 5, 50, and 95 of the damage indices by using IDA and PA. For these cases α values are 0.65, 0.78, and
0.67 for 5, 50, and 95 percentiles, respectively.

© ASCE 04019010-10 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


structural systems. Thus, the usefulness and success of the simpli-
fied method in practical applications also seems debatable. That
would be true if very different α values were obtained in different
applications, but the values found in this study for framed rein-
forced concrete structures have shown to be very stable, with minor
variations. For instance, Diaz-Alvarado et al. (2017) have recently
found α values of 0.71, 0.66, and 0.67, respectively, for low-rise,
midrise, and high-rise frame steel buildings, using earthquakes re-
corded in soft soils of Mexico City that have long durations. Thus,
such stability of the parameter α is a stronghold of the method. In
this paper, only the fundamentals of the method have been shown,
but ongoing work is being done in order to provide tabulated α
values for typical housing buildings including low-rise, midrise,
and high-rise, unreinforced masonry, reinforced concrete, and steel
buildings, typical of European cities and towns. Besides, α values
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

are being calibrated for seismic actions compatible with several re-
sponse spectra. Thus, damage curves might be obtained in a fast
and straightforward way for several buildings typologies and seis-
mic hazards. Moreover, the versatility of this new damage index
also allows providing adequate corrective factors for different kinds
of seismic actions and/or for different types of soils. The use of
these factors would improve the accuracy of the new damage func-
tions. Moreover, conservative α values can be taken for required or
predefined confidence levels. However, it also must be said that the
calibration of this parameter α for irregular or involved buildings
and structural systems would not be simple.
Fig. 15 highlights the fact that the proposed method allows
reproducing, in a simple and straightforward way, the results ob-
tained by using more sophisticated stochastic nonlinear dynamic
analysis. Moreover, the computational effort is much lower due to
the simple single degree of freedom approaches involved and also
to the easy computation of the performance point.
Finally, Fig. 16 shows a flow chart, depicting the primary steps Fig. 16. Flow chart depicting the computation of the capacity-based
involved in the computation of the new capacity-based damage in- damage index as a function of the PGA.
dices. Please recall at this point that 100 capacity curves (Counter 1)
have been already obtained and stored. Moreover, for each capacity
curve, 10 Sd-PGA curves (Counter 2) are also available; Sd holds
for the spectral displacement of the performance point. Then, there (2) the elastic response spectrum of the record, and (3) the crossing
is a curve for each earthquake record, each curve having 20 points capacity and response spectrum to get the performance point.
corresponding to 20 intensity levels; that is, 1,000 Sd-PGA curves Crossing capacity and response spectrum are a quick and straight-
have been also computed and stored previously in the MRSd-PGA forward process. Obtaining the response spectrum implies integrat-
matrix. These 1,000 curves have been shown already in Fig. 7. ing the equations of motion of a single degree of freedom linear
Then, for an assumed value of α, the steps to obtain the new damage system defined by means of the period and damping. The obtaining
index as a function of PGA are summarized as follows: of capacity spectrum is the most demanding part of the simplified
1. For each capacity curve (Counter 1), compute stiffness and en- method. Typically, the overall PA for the building analyzed in this
ergy functions according to Eqs. (6) and (7) and combine them paper, including the computation of the capacity curve and of the
according to the assumed α value; a damage index as a function performance point on a i7 3GHz processor, takes no more than
of the Sd is obtained. 240 s. Taking into account that obtaining a complete damage curve
2. for each earthquake record (Counter 2) load the corresponding requires estimating the performance point 20 times, but that this
Sd-PGA curve from the MRSd-PGA matrix. This way, a damage calculation does not require the computation of either the capacity
index as a function of the PGA is obtained (these curves for a α curve or of the response spectrum, the overall time for a single
value of 0.5 have been shown in Fig. 13). Check for the number damage curve by means of static analysis is no more than 400 s.
of records and for the number of buildings. If the maximum Otherwise, nonlinear dynamic analyses require solving the motion
number of records (10 in this case) has not been reached, then equations of a multiple degree of freedom nonlinear system at
go to Step 2 again. Otherwise, if the maximum number of struc- each point of the record, which typically is described by means
tures has not been obtained, go to Step 1; otherwise end. of several thousands of time-acceleration pairs. Assuming a 30 s
bracketed duration or about 3,000 time-acceleration pairs, typical
computation times for a single dynamic analysis of the same struc-
Discussion of the Proposed Approach ture, on an i7-3GHz processor, are about 120 s. Moreover, it must
take into account that the dynamic analysis must be done for each
From a numerical point of view, the primary advantage of the pro- one of the scaled records (20 in the present case) so that overall
posed method lies on the computational cost. In fact, given a struc- computation time may be about 2,400 s; that is about six times
tural model and a seismic action, defined by an earthquake record, the computational cost of the PA-based method. Note that this dif-
the static analysis requires computing (1) the capacity spectrum, ference between the numerical costs increases when the number of

© ASCE 04019010-11 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


seismic intensities involved in the incremental dynamic analy- that brittle structures submitted to impulsive seismic actions lead to
sis does. α values slightly higher than 0.8. On the contrary, ductile structures
Moreover, the fact that many computations can be done in an submitted to long duration intense signals produce lower α values,
independent manner for different seismic actions and for different thus indicating that a significant contribution to damage comes
random structures, makes the procedure parallelizable in an easy from energy dissipation. However, the range of variation of α is
way; although this is also true for nonlinear dynamic analyses. narrow; values in the range ∼0.65 to 0.9 have been generally found.
Notice that special care must be put into the design and program- This fact demonstrates the stability, and at the same time the
ming of the routine or function used to compute the performance versatility of the capacity-based damage index.
point by crossing capacity and demand spectra. Actually, the fact Because the purpose of this study is not related to any specific
that response spectra of actual records are noisy functions may com- site, the target spectrum has not been linked to probabilistic seismic
plicate the finding of the performance point, leading to a potential hazard analyses. Moreover, simplified methods have been used to
bottleneck of the procedure. However, smoothing the response spec- select and to scale strong ground motions to perform the compari-
tra improves the numerical calculation of the performance point. son between the results of the nonlinear static and dynamic analy-
ses. However, it is recognized that more realistic target spectra and
more sophisticated scaling methods should be investigated in order
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

Conclusions to analyze and discuss the results in specific sites where probabi-
listic seismic hazard analysis (PSHA) results are available.
The primary objective of the present study was to develop a sim- Finally, several advantages of the new method are summarized
plified method for obtaining probabilistic damage curves of struc- as follows:
tures based on the pushover analysis. As a testbed, the physical • The computational effort is much lower than the required by
damage of a framed reinforced concrete building has been assessed, IDA. To be precise, at least a reduction to one-sixth of the com-
taking into account that the input variables are random. Not only putational effort was observed when using the new procedure;
the strength of the concrete, the elastic modulus of the steel, and • If the seismic hazard changes, the dynamic analysis requires one
other related mechanical properties have been treated as random to perform a new complete nonlinear dynamic structural ana-
variables, but also the seismic action has been considered in a sto- lyses, while the new procedure only requires one to compute the
chastic way. An important conclusion is that the results presented in new performance points as the capacity curve simply depends
this paper show significant uncertainties in the seismic response on the structure; and
and expected damage when taking into account the randomness • The new method also allows one to take into account the un-
of the input variables. With the objective of improving the damage certainties of the structural behavior and those of the seismic
analysis, a new procedure has been presented. It is based on the actions.
stiffness degradation and on the energy loss. The results calculated
by using this new approach show a very good agreement with the
results obtained by means of dynamic analysis. Acknowledgments
One of the most relevant conclusions of this study is that when
the procedures discussed in this paper are used to evaluate the ex- This research has been partially funded by the Ministry of Econ-
pected seismic damage of a structure the parameters influencing omy and Competitiveness (MINECO) of the Spanish Government
the seismic damage curves of the structures must be considered and by the European Regional Development Fund (FEDER) of the
as random. Simplified deterministic procedures based on character- European Union (EU) through projects referenced as CGL2011-
istic values, usually may lead to conservative results. Nevertheless, 23621 and CGL2015-65913 -P (MINECO/FEDER, UE).
some abridged assumptions on the definition of the seismic actions
and on the estimation of the seismic damage states and thresholds
can also underestimate the real damage that can occur in a structure. References
A structural typology, which can be classified as a low- to mid-
rise symmetric reinforced concrete framed building was studied. Abrahamson, N. A., P. G. Somerville, and C. A. Cornell. 1991. “Uncer-
The results are compared, from a probabilistic perspective, with the tainty in numerical ground motion predictions.” In Proc., 4th US
damage curves calculated with IDA, in terms of several percentiles. National Conf. of Earthquake Engineering, 407–416. Berkeley, CA:
The results clearly show the good fit of the damage curves calcu- EERI.
lated with the new method. However, so far the proposed method Ambraseys, N., P. Smit, R. Sigbjornsson, P. Suhadolc, and B. Margaris.
has been applied only to simplified symmetric models, where the 2004. “Internet-site for European strong-motion data.” European Com-
first mode of vibration dominates the seismic response and for mission, Research-Directorate General, Environment and Climate
which 2D models are enough to capture the structural response. For Programme. Accessed May 17, 2017. http://www.isesd.hi.is/ESD
tall, irregular, or asymmetric buildings further considerations should _Local/frameset.htm.
Arroyo, D., and M. Ordaz. 2011. “On the forecasting of ground-motion
be made by taking into account higher mode effects in elevation
parameters for probabilistic seismic hazard analysis.” Earthquake
(Chopra et al. 2004; Poursha et al. 2009), in plan (Chopra and Goel
Spectra 27 (1): 1–21. https://doi.org/10.1193/1.3525379.
2004; Bento et al. 2010; Bhatt and Bento 2011), or in both plan and
ATC (Applied Technology Council). 1996. Seismic evaluation and retrofit
elevation (Kreslin and Fajfar 2012; Reyes and Chopra 2010, 2011; of concrete buildings. ATC-40. Redwood City, CA: ATC.
Fujii 2011), among other aspects. In these cases, the influence of Barbat, A. H., M. L. Carreño, O. D. Cardona, and M. C. Marulanda. 2011.
higher modes should be reviewed and considered in order to en- “Evaluación holística del riesgo sísmico en zonas urbanas.”
hance the scope of the proposed method. Revista internacional de métodos numéricos para cálculo y diseño
A parameter α was used to weight the contribution to the global en ingeniería 27 (1): 3–27.
damage due to the ductility and the dissipated energy. This param- Bento, R., C. Bhatt, and R. Pinho. 2010. “Using nonlinear static proce-
eter depends on the structure, but also on the seismic actions. Prob- dures for seismic assessment of the 3D irregular SPEAR building.”
ably this is a strong point of the capacity-based damage index Earthquake Struct. 1 (2): 177–195. https://doi.org/10.12989/eas.2010
because of its versatility and robustness. Ongoing research indicate .1.2.177.

© ASCE 04019010-12 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Bhatt, C., and R. Bento. 2011. “Assessing the seismic response of existing Fragiadakis, M., and D. Vamvatsikos. 2010. “Fast performance uncertainty
RC buildings using the extended N2 method.” Bull. Earthquake Eng. estimation via pushover and approximate IDA.” Earthquake Eng.
9 (4): 1183–1201. https://doi.org/10.1007/s10518-011-9252-8. Struct. Dyn. 39 (6): 683–703.
Bommer, J. J., and H. Crowley. 2006. “The influence of ground motion Franchin, P., P. Pinto, and R. Pathmanathan. 2010. “Confidence factor?”
variability in earthquake loss modelling.” Bull. Earthquake Eng. J. Earthquake Eng. 14 (7): 989–1007. https://doi.org/10.1080
4 (3): 231–248. https://doi.org/10.1007/s10518-006-9008-z. /13632460903527948.
Bommer, J. J., P. J. Stafford, J. E. Alarcón, and S. Akkar. 2007. “The in- Freeman, S. A. 1998. “Development and use of capacity spectrum method.”
fluence of magnitude range on empirical ground-motion prediction.” In Proc., 6th US National Conf. of Earthquake Engineering. Seattle:
Bull. Seismol. Soc. Am. 97 (6): 2152–2170. https://doi.org/10.1785 EERI.
/0120070081. Freeman, S. A., J. P. Nicoletti, and J. V. Tyrell. 1975. “Evaluations of
Brozovič, M., and M. Dolšek. 2014. “Envelope-based pushover analysis existing buildings for seismic risk: A case study of Puget sound naval
procedure for the approximate seismic response analysis of buildings.” shipyard, Bremerton, Washington.” In Proc. 1st US National Conf. of
Earthquake Eng. Struct. Dyn. 43 (1): 77–96. https://doi.org/10.1002 Earthquake Engineering, 113–122. Berkeley, CA: EERI.
/eqe.2333. Fujii, K. 2011. “Nonlinear static procedure for multi-story asymmetric
building considering bi-directional excitation.” J. Earthquake Eng.
Carr, A. J. 2000. Ruaumoko-inelastic dynamic analysis program.
15 (2): 245–273. https://doi.org/10.1080/13632461003702902.
Christchurch, New Zealand: Dept. of Civil Engineering, Univ. of
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

Hancock, J., J. J. Bommer, and P. J. Sttaford. 2008. “Numbers of scaled and


Canterbury.
matched accelerograms required for inelastic dynamic analyses.” Earth-
Celarec, D., and M. Dolšek. 2013. “The impact of modelling uncertain-
quake Eng. Struct. Dyn. 37 (14): 1585–1607. https://doi.org/10.1002
ties on the seismic performance assessment of reinforced concrete
/eqe.827.
frame buildings.” Eng. Struct. 52 (Jul): 340–354. https://doi.org/10
Kreslin, M., and P. Fajfar. 2012. “The extended N2 method considering
.1016/j.engstruct.2013.02.036. higher mode effects in both plan and elevation.” Bull. Earthquake
CEN (European committee for standardization). 2004a. Design of concrete Eng. 10 (2): 695–715. https://doi.org/10.1007/s10518-011-9319-6.
structures. Part 1: General-common rules for building and civil engi- Lagomarsino, S., and S. Giovinazzi. 2006. “Macroseismic and mechanical
neering structures. Eurocode 2. Brussels, Belgium: CEN. models for the vulnerability and damage assessment of current build-
CEN (European committee for standardization). 2004b. Design of struc- ings.” Bull. Earthquake Eng. 4 (4): 415–443. https://doi.org/10.1007
tures for earthquake resistance. Part 1: General rules, seismic actions /s10518-006-9024-z.
and rules for building. Eurocode 8. Brussels, Belgium: CEN. Lantada, N., J. Irrizari, A. H. Barbat, X. Goula, A. Roca, T. Susagna, and
Chopra, A. K., and R. K. Goel. 1999. Capacity-demand-diagram methods L. G. Pujades. 2010. “Seismic hazard and risk scenarios for Barcelona,
for estimating seismic deformation of inelastic structures: SDOF sys- Spain, using the risk-UE vulnerability index method.” Bull. Earthquake
tems. PEER Rep. No. 1999/02. Berkeley, CA: Pacific Earthquake Eng. 8 (2): 201–229. https://doi.org/10.1007/s10518-009-9148-z.
Engineering Research Center, Univ. of California. Lin, T., C. Haselton, and J. W. Baker. 2013a. “Conditional spectrum-based
Chopra, A. K., and R. K. Goel. 2002. “A modal pushover analysis pro- ground motion selection. Part I: Hazard consistency for risk-based as-
cedure for estimating seismic demand buildings.” Earthquake Eng. sessments.” Earthquake Eng. Struct. Dyn. 42 (12): 1847–1865. https://
Struct. Dyn. 31 (3): 561–582. https://doi.org/10.1002/eqe.144. doi.org/10.1002/eqe.2301.
Chopra, A. K., and R. K. Goel. 2004. “A modal pushover analysis pro- Lin, T., C. Haselton, and J. W. Baker. 2013b. “Conditional spectrum-
cedure to estimate seismic demand for unsymmetric-plan buildings.” based ground motion selection. Part II: Intensity-based assessments
Earthquake Eng. Struct. Dyn. 33 (8): 903–927. https://doi.org/10 and evaluation of alternative target spectra.” Earthquake Eng. Struct.
.1002/eqe.380. Dyn. 42 (12): 1867–1884. https://doi.org/10.1002/eqe.2303.
Chopra, A. K., R. K. Goel, and C. Chintanapakdee. 2004. “Evaluation of a Melchers, R. E. 1999. Structural reliability analysis and prediction.
modified MPA procedure assuming higher modes as elastic to estimate New York: Wiley.
seismic demands.” Earthquake Spectra 20 (3): 757–778. https://doi.org Otani, S. 1974. “Inelastic analysis of RC frame structures.” J. Struct. Div.
/10.1193/1.1775237. 100 (ST7): 1433–1449.
Crowley, H., J. J. Bommer, R. Pinho, and J. F. Bird. 2005. “The impact of Park, Y. J., and A. H. S. Ang. 1985. “Mechanistic seismic damage model
epistemic uncertainty on an earthquake loss model.” Earthquake Eng. for reinforced concrete.” J. Struct. Eng. 111 (4): 722–739. https://doi
Struct. Dyn. 34 (14): 1653–1685. https://doi.org/10.1002/eqe.498. .org/10.1061/(ASCE)0733-9445(1985)111:4(722).
Diaz-Alvarado, S. A., L. G. Pujades, A. H. Barbat, D. A. Hidalgo-Leiva, Park, Y. J., A. H. S. Ang, and Y. K. Wen. 1985. “Seismic damage analysis
and Y. F. Vargas. 2017. “Capacity, damage and fragility models for steel of reinforced concrete buildings.” J. Struct. Eng. 111 (4): 740–757.
buildings. A probabilistic approach.” Bull. Earthquake Eng. 16 (3): https://doi.org/10.1061/(ASCE)0733-9445(1985)111:4(740).
1209–1243. Poursha, M., F. Khoshnoudian, and A. S. Moghadam. 2009. “A consecutive
modal pushover procedure for estimating the seismic demands of tall
Eads, L., E. Miranda, H. Krawinkler, and D. Lignos. 2013. “An efficient
buildings.” Eng. Struct. 31 (2): 591–599. https://doi.org/10.1016/j
method for estimating the collapse risk of structures in seismic regions.”
.engstruct.2008.10.009.
Earthquake Eng. Struct. Dyn. 42 (1): 25–41. https://doi.org/10.1002
Pujades, L. G., Y. F. Vargas, A. H. Barbat, and J. R. González-Drigo. 2015.
/eqe.2191.
“Parametric model for capacity curves.” Bull. Earthquake Eng. 13 (5):
Egozcue, J. J., A. H. Barbat, J. A. Canas, J. Miquel, and E. Banda. 1991. “A
1347–1376. https://doi.org/10.1007/s10518-014-9670-5.
method to estimate occurrence probabilities in low seismic activity
Reyes, J. C., and A. K. Chopra. 2010. “Three-dimensional modal pushover
regions.” Earthquake Eng. Struct. Dyn. 20 (1): 43–60. https://doi.org/10 analysis of building subjected to two components of ground motions,
.1002/eqe.4290200105. including its evaluation for tall buildings.” Earthquake Eng. Struct.
Fajfar, P. 1999. “Capacity spectrum based on inelastic demand spectra.” Dyn. 40 (7): 789–806. https://doi.org/10.1002/eqe.1060.
Earthquake Eng. Struct. Dyn. 28 (9): 979–993. https://doi.org/10 Reyes, J. C., and A. K. Chopra. 2011. “Evaluation of three-dimensional
.1002/(SICI)1096-9845(199909)28:9<979::AID-EQE850>3.0.CO;2-1. modal pushover analysis for unsymmetric-plan buildings subjected to
Fajfar, P., and P. Gaspersic. 1996. “The N2 method for the seismic two components of ground motion.” Earthquake Eng. Struct. Dyn.
damage analysis of RC buildings.” Earthquake Eng. Struct. Dyn. 25 (1): 40 (13): 1475–1494. https://doi.org/10.1002/eqe.1100.
31–46. https://doi.org/10.1002/(SICI)1096-9845(199601)25:1<31::AID Satyarno, I. 2000. “Adaptive pushover analysis for the seismic assessment
-EQE534>3.0.CO;2-V. of older reinforced concrete buildings.” Ph.D. thesis, Dept. of Civil
FEMA. 1999. HAZUS earthquake loss estimation method. Washington, Engineering, Univ. of Canterbury.
DC: FEMA. Vamvatsikos, D., and C. A. Cornell. 2002. “The incremental dynamic
FEMA. 2005. Improvement of nonlinear static seismic procedures. FEMA analysis.” Earthquake Eng. Struct. Dyn. 31 (3): 491–514. https://doi
440. Washington, DC: FEMA. .org/10.1002/eqe.141.

© ASCE 04019010-13 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010


Vamvatsikos, D., and M. Fragiadakis. 2010. “Incremental dynamic analysis y diseño en ingeniería 29 (2): 63–78. https://doi.org/10.1016/j.rimni
for estimating seismic performance sensitivity and uncertainty.” .2013.04.003.
Earthquake Eng. Struct. Dyn. 39 (2): 141–163. Vargas, Y. F., L. G. Pujades, A. H. Barbat, and J. E. Hurtado. 2015. “Prob-
Vargas, Y. F., L. G. Pujades, A. H. Barbat, and J. E. Hurtado. 2012. abilistic seismic damage assessment of RC buildings based on nonlinear
“Incremental dynamic analysis and pushover analysis of buildings.” dynamic analysis.” Open Civ. Eng. J. 9 (1): 344–350. https://doi.org/10
In Vol. 2 of Computational methods in stochastic dynamics. New York: .2174/1874149501509010344.
Springer. Vargas, Y. F., L. G. Pujades, A. H. Barbat, J. E. Hurtado, S. A. Diaz-
Vargas, Y. F., L. G. Pujades, A. H. Barbat, and J. E. Hurtado. 2013a. Alvarado, and D. A. Hidalgo-Leiva. 2017. “Probabilistic seismic
“Capacity, fragility and damage in reinforced concrete buildings: damage assessment of reinforced concrete buildings considering direc-
A probabilistic approach.” Bull. Earthquake Eng. 11 (6): 2007–2032. tionality effects.” Struct. Infrastruct. Eng. 14 (6): 817–829. https://doi
https://doi.org/10.1007/s10518-013-9468-x. .org/10.1080/15732479.2017.1385089.
Vargas, Y. F., L. G. Pujades, A. H. Barbat, and J. E. Hurtado. 2013b. Williams, S. M., and R. G. Sexsmith. 1995. “Seismic damage indices
“Evaluación probabilista de la capacidad, fragilidad y daño sísmico for concrete structures: A state of the art review.” Earthquake Spectra
en edificios de hormigón armado.” Métodos numéricos para cálculo 11 (2): 319–349. https://doi.org/10.1193/1.1585817.
Downloaded from ascelibrary.org by Sergio Barreiro on 07/20/19. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04019010-14 J. Struct. Eng.

J. Struct. Eng., 2019, 145(4): 04019010

You might also like