Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Bulletin of Earthquake Engineering

https://doi.org/10.1007/s10518-020-00885-1

S.I. : RECENT ADVANCES IN SEISMIC FRAGILITY AND VULNERABILITY

Development of a fragility and vulnerability model for global


seismic risk analyses

Luís Martins1   · Vítor Silva1,2

Received: 13 December 2019 / Accepted: 1 June 2020


© Springer Nature B.V. 2020

Abstract
Seismic fragility and vulnerability assessment is an essential step in the evaluation of prob-
abilistic seismic risk. Ideally, models developed and calibrated for the building portfolio
of interest would be readily available. However, the lack of damage data and insufficient
analytical studies lead to a paucity of fragility and vulnerability models, in particular in
the developing world. This study describes the development of an analytical fragility and
vulnerability model covering the most common building classes at the global scale. Nearly
five hundred functions were developed to cover the majority of combinations of construc-
tion material, height, lateral load resisting system and seismic design level. The fragility
and vulnerability were derived using nonlinear time-history analyses on equivalent single-
degree-of-freedom oscillators and a large set of ground motion records representing several
tectonic environments. The resulting fragility and vulnerability functions were validated
through a series of tests which include the calculation of the average annual loss ratio for a
number of locations, the comparison of probabilities of collapse across all building classes,
and the repetition of past seismic events. The set of vulnerability functions was used for the
assessment of economic losses due to earthquakes as part of the global seismic risk model
supported by the Global Earthquake Model Foundation.

Keywords  Fragility assessment · Vulnerability assessment · Seismic risk assessment ·


Global database

1 Introduction

In the last century earthquakes have caused over one million fatalities and nearly one
trillion USD in economic losses (Elnashai and Di Sarno 2008; Sen 2009). Informal con-
struction or structures built following inadequate seismic regulations are responsible for
the majority of the human and economic losses (Bull 2013; Calvi 2010; Gautam and

* Luís Martins
luis.martins@globalquakemodel.org
1
Global Earthquake Model Foundation, Pavia, Italy
2
Faculty of Science and Technology, University Fernando Pessoa, Porto, Portugal

13
Vol.:(0123456789)
Bulletin of Earthquake Engineering

Fig. 1  Number of fragility or vulnerability functions per country in the OpenQuake-platform

Chaulagain 2016; So 2016). It is thus relevant to evaluate the seismic vulnerability of the
building stock in order to improve our understanding of the geographic distribution of
seismic risk, and to support the development and implementation of risk mitigation strate-
gies. These can include the development of retrofitting campaigns (Ferreira et  al. 2016),
the enforcement of modern design regulations (Burby et al. 1998) or the development of
catastrophe insurance polls to transfer the risk from the public sector to the international
reinsurance market (Bommer et al. 2002).
Fragility curves describe the probability of exceeding a set of damage states conditional
to a ground shaking intensity, and are fundamental for the assessment of damage in earth-
quake scenarios (e.g. Acevedo et al. 2017). These functions can be converted into vulner-
ability functions using a damage-to-loss model (i.e. relation between a damage state and
the corresponding fraction of loss), leading to a distribution of probability of loss ratio
conditional on a set of ground shaking intensities.
As part of the Global Earthquake Model (GEM) initiative, hundreds of empirical and
analytical fragility and vulnerability functions were collected (D’Ayala et  al. 2015; Ros-
setto et  al. 2014; Yepes-Estrada et  al. 2016), and made publicly available through the
OpenQuake-platform (https​://platf​orm.openq​uake.org/vulne​rabil​ity). Despite the notable
increase in the availability of fragility and vulnerability functions in the last decade, there
are still several regions in the world where such models do not exist, or are simply not pub-
licly available, as illustrated in Fig. 1.
Fragility models suitable for large-scale seismic risk assessment have been proposed for
few regions of the world either as a result of independent studies, regional programmes,
or government-funded research projects. For example, the well-known HAZUS software
package supported by the Federal Emergency and Management Agency of the United
States (FEMA 2014) includes more than 100 fragility models for the most common build-
ing classes in the United States. In Europe, several projects funded by the European Com-
mission have produced fragility or vulnerability functions for some types of construction.
These include the RISK-UE project (2001–2004) (Mouroux and Brun 2006), the LESS-
LOSS project (2004–2007) (Calvi and Pinho 2004), and more recently, the SERA project
(2017–2020) (Crowley et  al. 2018). In South America, Villar-Vega et  al. (2017) in col-
laboration with local experts proposed a set of fragility functions for 54 building classes.

13
Bulletin of Earthquake Engineering

For the African continent, the GEM Foundation led an initiative to characterize the most
common building classes, and develop suitable vulnerability functions (https​://www.globa​
lquak​emode​l.org/afric​a-model​-relea​se). At the national level, there are a number of studies
that generated fragility functions for the majority of the building classes in the country.
Some examples include Erberik (2008a, b) and Erdik et al. (2003) for Turkey, Borzi et al.
(2008a, b) for Italy, Salgado-Gálvez et  al. (2013) for Colombia, Silva et  al. (2014a) for
Portugal, and Motamed et al. (2018) for Iran. All of these models are publicly accessible
through the OpenQuake-platform.
Despite the proven value of these existing models in the assessment of seismic risk at
the national or regional scale (e.g. Crowley et al. 2004; Jaiswal and Wald 2011; Silva et al.
2014b), there are a number of reasons that support the development of a global model:

i. As illustrated in Fig. 1, there are still a large number of countries (and consequently
building classes) that have not been covered by any vulnerability study.
ii. The employment of distinct vulnerability assessment methodologies can lead to sig-
nificantly different risk results (e.g. Strasser et al. 2008; Silva et al. 2013), thus pre-
venting an unbiased comparison of the risk across regions or countries that considered
different methodologies.
iii. Some of the existing models have used macroseismic intensity for the definition of
the fragility and/or vulnerability functions, which hampers their combinations with
modern probabilistic seismic hazard analysis (PSHA) models. In the vast majority of
the cases, modern seismic hazard assessment studies use ground motion parameters
(e.g. peak ground acceleration, spectral acceleration).
iv. Recent studies have demonstrated the value in exploring more efficient intensity meas-
ures for the definition of fragility functions, such as average spectral acceleration (e.g.
Kohrangi et al. 2017). To the best of the authors knowledge, such intensity measures
have yet to be considered in vulnerability studies.
v. Some of the existing fragility and vulnerability models have been developed for spe-
cific buildings, and not necessarily for a given building class. The disregard of the
building-to-building variability will lead to an underestimation of the range of pos-
sible losses (e.g. Silva et al. 2019).
vi. Ideally, fragility models and damage-to-loss models should not only be available but
also be compatible (e.g. Martins et al. 2016). However, in some cases the definition
of the damage states in the fragility model might not be directly comparable with the
damage states in the damage-to-loss model (e.g. Borzi et al. 2008a; Di Pasquale and
Goretti 2001).
vii. Finally, only a few of the existing fragility and/or vulnerability models have been actu-
ally employed in the assessment of probabilistic seismic risk or estimation of losses
due to earthquake scenarios (e.g. Villar-Vega et al. 2017). This lack of validation and
verification might lead to unrealistic seismic risk estimates.

Having identified the need to develop a comprehensive global database of fragility mod-
els and considering the results of a worldwide scale survey to define the most common
building classes, analytical fragility and vulnerability models were developed for nearly
500 building classes. The vulnerability models were derived from nonlinear dynamic anal-
yses performed on equivalent single-degree-of-freedom (SDOF) oscillators considering the
building-to-building and the record-to-record variability. The probability of exceeding a set
of damage states was estimated through cloud analysis (Jalayer et al. 2015) and considering

13
Bulletin of Earthquake Engineering

the uncertainty in the definition of the damage criterion. The conversion of the fragility
functions into vulnerability functions was performed considering existing empirical and
numerical damage-to-loss models, and propagating the variability in the loss ratio per dam-
age state. To ensure that the models produced reasonable seismic risk estimates, determin-
istic and probabilistic seismic risk metrics were explored. All of the parameters used in the
fragility and vulnerability derivation process are included in a public GitHub repository.1
This set of global fragility and vulnerability functions were used for the development of
the global seismic risk map2 (Silva et al. 2019) released by the Global Earthquake Model
(GEM) Foundation and its partners in December 2018.

2 Identification of common building classes

To assist in the development of a comprehensive catalogue of the most representative


building classes worldwide, a survey targeted at experts in the field of structural and earth-
quake engineering was conducted by the GEM Foundation. A web-based application that
allows users to classify a particular building stock based on the most significant structural
parameters (e.g. building material, lateral load resisting system, height, level of ductility,
irregularities) was developed. For each country, users are also asked to quantify the relative
frequency of a given building class in both rural or urban areas. In addition, the applica-
tion allows the classification of buildings according to their use (i.e. residential, commer-
cial, industrial, healthcare, governmental and educational). This application is accessible
through the OpenQuake-platform at https​://platf​orm.openq​uake.org/build​ing-class​.
At the time of writing, nearly 600 entries were collected covering over 70 countries.
Such number is obviously far from being exhaustive, and is statistically insufficient to state
that the most relevant building classes at the global scale were covered. To improve the list
of possible building classes, four other sources of information were considered:

i. Past regional and national projects (e.g. SARA, SERA) which featured activities
related with the classification of the building stock, as described in the previous sec-
tion.
ii. Technical reports from the World Housing Encyclopaedia (www.world-​ housin​ g.net/),
which have a greater coverage for vulnerable structures in less developed nations.
iii. The outcomes from the PAGER-WHE project (Jaiswal et al. 2010), which collected
information from more than 30 countries in a standardized format.
iv. Regional workshops led by GEM with sessions dedicated to the identification of
the most relevant building class (East Sub-Saharan Africa—Addis Ababa; Europe—
Pavia; Central Asia—Bishkek; India—Gandhinagar; South Asia—Kathmandu; South
America—Lima and Medellin; Central America and the Caribbean—San Jose and
Santo Domingo; South-East Asia—Bandung).

The current list of building classes considered buildings divided according to (i) the
construction material, (ii) the lateral load resisting system, (iii) ductility level and (iv)
height. For what concerns the material of construction, users have identified the following

1
  https​://githu​b.com/lmart​ins88​/globa​l_fragi​lity_vulne​rabil​ity.
2
  Global Seismic Risk Map: https​://maps.openq​uake.org/map/globa​l-seism​ic-risk-map/.

13
Bulletin of Earthquake Engineering

Fig. 2  Distribution of building classes for Peru according to material and height

main materials: masonry (unreinforced (MUR), reinforced (MR) or confined (MCF)), rein-
forced concrete (CR), adobe/earthen (MUR-ADO), timber/wood (W), steel (S) and com-
posite (SCR) (i.e. combination of reinforced concrete with steel).
Regarding the lateral load-resisting system, the results indicate four main lateral load
resisting systems: moment resisting bare frames (LFM), infilled frames (LFINF), wall
systems (LWAL) and dual systems (LDUAL). For each lateral load-resisting system three
levels of ductility were considered (i.e. low, moderate and high ductility) to reflect dif-
ferent levels of seismic design. Structures designed with very limited seismic provisions
(i.e. mostly informal construction or structures built prior to 1960) or structures in regions
where earthquakes are not the dominant load combination (i.e. low seismic hazard regions)
were classified with a low ductility (DUL). Structures that incorporated some degree of
seismic design but with regulations deemed less rigorous than modern codes were grouped
under the moderate ductility (DUM) category (i.e. generally buildings from late 70′s to
early 90′s, or built recently but in areas with low seismic hazard). Structures designed
according to modern design regulations (commonly considered to have been introduced
between late 90′s and early 2000′s) were classified under high ductility (DUH). One excep-
tion to this classification system was introduced for the unreinforced masonry structures
or earthen construction, which due to the expected poor performance, were considered as
non-ductile (DNO).
Finally, for the characterization of the height of the buildings (and for the sake of sim-
plicity) users classified the building stock according to four main categories: low-, mid-,
high-rise and tall structures. These categories were then further discretized in actual num-
ber of storeys for the development of the fragility and vulnerability functions. Such deci-
sion was motivated by the fact that the expected capacity of some typologies (e.g. masonry)
revealed to be reasonably sensitive to the number of storeys (e.g. Borzi et al. 2008a). An
example of a building class survey for Peru is summarized in Fig. 2. For Peru the survey
has identified a total of 115 building classes mostly consisting of low-rise masonry classes.

13
Bulletin of Earthquake Engineering

The current list of building classes comprises almost 500 typologies, as listed in the
public GitHub repository. However, it should be noted that the framework presented herein
aims at being a dynamic platform, which will be constantly updated as new information
regarding existing building classes is collected. The data and tools used in this study are
available through public repositosries to the scientific community, and the derivation
framework is further discussed in the following sections.

3 Fragility and vulnerability assessment

In the last decades, dozens of analytical methodologies have been proposed for the devel-
opment of fragility functions, spanning from complex analysis of three-dimensional
numerical models (e.g. Martins et  al. 2016; Ulrich et  al. 2014), to more simplified and
expedite methods (e.g. Vamvatsikos and Allin Cornell 2006; Silva et al. 2013). A compre-
hensive review of existing fragility derivation methodologies can be found in D’Ayala et al.
(2015). For this study, given the need to consider a large number of building classes, while
still accounting for a wide range of uncertainties, a decision was made to employ nonlin-
ear time history analysis on equivalent single-degree-of-freedom (SDOF) oscillators (e.g.
Dolšek 2012; Villar-Vega et al. 2017).
These analyses were performed using the Risk Modellers Toolkit (RMTK) supported by
the GEM Foundation (Silva et al. 2015). The RMTK is an open source platform targeted
at earthquake engineers, and its code is publicly accessible through a GitHub repository
(https​://githu​b.com/GEMSc​ience​Tools​/rmtk). The toolkit is fully integrated with the finite
element software OpenSees (McKenna et al. 2000) and is available for all main operating
systems. The following sub-sections provide a description of the components required for
the derivation of the fragility and vulnerability functions.

3.1 Ground motion record selection

This study aims at developing a global database of fragility and vulnerability functions,
and therefore a more general approach to the selection of the ground motion records
was explored, as opposed to methodologies that optimise the selection for a specific site
(e.g. conditional spectrum method—Baker 2011). In this context, records were selected
according to the tectonic environment, magnitude and distance. The accelerograms were
collected from several ground motion databases worldwide (e.g. Pacific Earthquake Engi-
neering Research (PEER) NGA-West, Chilean Geological Institute, Colombian Geological
Service, Universidad Nacional Autónoma de México—Engineering Instituting, European
Strong Motion Database) for active shallow and subduction tectonic environments. Over
3500 records with a minimum peak ground acceleration (PGA) of 0.05 g were collected.
Figure 3 illustrates the locations of the recording stations of these records.
From the initial catalogue of 3500 ground motion records, 300 signals per tectonic
regime and intensity measure (IM) were selected for the nonlinear dynamic analyses.
Four IMs were considered in order to account for the dynamic properties of the differ-
ent building classes: (i) PGA for stiff/low-rise structures, (ii) Spectral acceleration at 0.3
seconds (SA0.3 s) for some low and mid-rise structures, (iii) Spectral acceleration at 0.6
seconds (SA0.6  s) for mid-rise structures, and finally (iv) Spectral acceleration at 1.0
seconds (SA1.0  s) for flexible/high-rise structures. For each IM, ten bins of acceleration
between 0.10 g and 2.0 g were considered, and 30 records were randomly selected from

13
Bulletin of Earthquake Engineering

Fig. 3  Location of the recording stations of the selected ground motion records

the catalogue for each bin. When 30 records were not available for a particular bin (e.g. in
the high intensity range), the records were scaled up to a maximum of 2.0 times (e.g. Wat-
son-Lamprey and Abrahamson 2006) until the minimum number of records was achieved.
This number of records per bin was assumed to be enough to achieve statistically sufficient
results in the structural response of the buildings (e.g. Sousa et al. 2016).
It is important to understand the influence of the tectonic environment in the resulting
fragility analysis. As discussed by Bommer et al. (2009), records from subduction environ-
ments tend to have a longer duration compared to shallow crustal events. Liel et al. (2015)
has commented that failing to consider the characteristics of subduction events can lead to
an underestimation of seismic risk in regions whose hazard is dominated by this type of
events. However, as the purpose of this study is to provide sets of fragility and vulnerability
functions for global risk analyses, records from both shallow and subduction regions were
used. The derivation of functions for each tectonic regime and their explicit incorporation
in probabilistic seismic risk analysis is discussed in the concluding remarks.

3.2 Definition of SDOF oscillators per building class

The methodology for fragility assessment followed in this manuscript relies on nonlinear
time-history analyses performed on equivalent SDOF oscillators. The hysteresis behav-
iour of each SDOF oscillator (one per building class) was defined based on the associated
capacity curve, expressed in terms of spectral acceleration (Sa), versus spectral displace-
ment (Sd). For the building classes where no significant reduction in the base shear capac-
ity due to damage accumulation was expected (e.g. steel frames, timber and composite
structures) the capacity was assumed to follow a trilinear elasto-plastic model (see Fig. 4).
The curves were defined based on the yielding (­ Sdy) and ultimate (­ Sdult) displacements and
the elastic ­(T1) and yielding ­(Ty) periods. When a formula to directly calculate the yielding
period was not available, it was assumed that ­Ty is equal to 1.5 times ­T1 (e.g. Calvi et al.
2006; Crowley and Pinho 2006), as indicated in Table 1.
Building classes for which a significant reduction in the base shear capacity is expected
due to the accumulation of damage in masonry panels (e.g. confined masonry or rein-
forced concrete with infills), e.g. Dolšek and Fajfar 2008; Riahi et al. 2009) were modelled

13
Bulletin of Earthquake Engineering

Fig. 4  Example of generic capacity curve

assuming a quadrilinear capacity curve (as depicted in Fig.  5). The reduction in the dis-
placement capacity due to the additional stiffness from the masonry walls was modelled
using the reduction factors proposed by Bal et al. (2010a). For these building classes a reg-
ular distribution of the infill walls was assumed, leading to a similar post-peak behaviour to
the one presented in Villar-Vega et al. (2017). Building classes with this type of backbone
curve follow a similar approach as in Fig. 4 to compute the first two points, with the end
of the elastic range being defined at 0.5 ­Say and ­T1 and the peak being computed from the
yield displacement and T ­ y. The third point of the capacity curve coincides with the yield
displacement of the equivalent bare frame class. Finally, the last capacity point is simply
taken from the ultimate displacement.
In Acceleration Displacement Response Spectrum (ADRS) format, the mean capacity
is estimated from the roof displacement (δroof) of the multi-degree-of-freedom structure
considering the relationships in Eqs. (1) and (2), where (T) is the period of vibration and
(Г) the first mode participation factor. Table 1 lists the parameters used to define the capac-
ity curves of each building class resulting of the compilation of data coming from research
studies and experimental campaigns. The classification of each building class followed the
GEM taxonomy (Brzev et al. 2013) guidelines revised according to (Silva et al. 2017).
𝛿roof
Sd = (1)
𝛤

( )2
2𝜋
Sa = Sd (2)
T

3.3 Damage allocation, fragility and vulnerability assessment methodologies

Four damage states have been considered ranging from slight (DS1) to complete damage
(DS4). To account for damage initiation in non-structural elements (e.g. infill walls), slight

13
Table 1  Structural and dynamic parameters used to define the capacity curves
Building class Ty [s] No. of storeys Inter-sto- Yield drift [%] Ult. drift [%] Γ References
rey height
[m]

MUR + ADO/LWAL + DNO Ty = 0.066H 0.9 1–3 2.6 0.12 0.55 1.4 Tarque et al. (2012), Preciado et al. (2020),
Karanikoloudis and Lourenço (2018), Tarque
et al. (2010)
MUR/LWAL + DNO Ty = 0.062H 0.9 1–5 2.8 0.14 0.60 1.4 Snoj and Dolšek (2017), Borzi et al. (2008a), b),
MUR + STRUB/LWAL + DNO Ty = 0.062H 0.9 1–5 2.8 0.12 0.51 1.4 Villar-Vega et al. (2017), Calvi (1999), Erberik
Bulletin of Earthquake Engineering

(2008b)
MUR + STDRE/LWAL + DNO Ty = 0.062H 0.9 1–5 2.8 0.13 0.54 1.4
MUR + CL99/LWAL + DNO Ty = 0.065H 0.9 1–5 2.8 0.13 0.57 1.4
MUR + CB99/LWAL + DNO Ty = 0.065H 0.9 1–5 2.8 0.16 0.69 1.4
MR/LWAL + DUL Ty = 0.060H 0.9 1–5 2.8 0.17 0.80 1.5 Lotfy et al. (2019), Hamzeh et al. (2018),Mur-
MR/LWAL + DUM Ty = 0.058H 0.9 1–5 2.8 0.20 0.90 1.5 cia-Delso and Shing (2012)
MR/LWAL + DUH Ty = 0.056H 0.9 1–5 2.8 0.22 1.00 1.5
MCF/LWAL + DUL Ty = 0.044H 1–6 2.8 0.19 0.90 1.5 Ahmad et al. (2010), Lovon et al. (2018),
MCF/LWAL + DUM Ty = 0.042H 1–6 2.8 0.24 1.00 1.5 Tomaževič and Klemenc (1997), Alcocer et al.
(2004), Tomaževič and Gams (2012), García
MCF/LWAL + DUH Ty = 0.04H 1–6 2.8 0.28 1.10 1.5
and Degrande (2017), Riahi et al. (2009),
Yañez et al. (2004)
CR/LFM + DUL Ty = 0.10H 1–12 2.8 0.73 3.00 1.4 Crowley and Pinho (2004, 2006, 2010), Ghoba-
CR/LFM + DUM Ty = 0.10H 1–12 2.8 0.75 3.20 1.4 rah (2004), Martins et al. (2018), Rossetto and
Elnashai (2003)
CR/LFM + DUH Ty = 0.10H 1–12 2.8 0.77 3.40 1.4

13

Table 1  (continued)
Building class Ty [s] No. of storeys Inter-sto- Yield drift [%] Ult. drift [%] Γ References
rey height

13
[m]

CR/LFINF + DUL Ty = 0.046H 1–12 2.8 0.15 1.20 1.33 Crowley and Pinho (2004, 2006, 2010), Grubišić
et al. (2013), Del Gaudio et al. (2019), Hoult
CR/LFINF + DUM Ty = 0.044H 1–12 2.8 0.20 1.35 1.33 et al. (2019)
CR/LFINF + DUH Ty = 0.042H 1–12 2.8 0.25 1.50 1.33
CR/LWAL + DUL Ty = 0.088N 1–12 2.8 0.17 1.10 1.4
CR/LWAL + DUM Ty = 0.083N 1–12 2.8 0.22 1.20 1.4
CR/LWAL + DUH Ty = 0.078N 1–12 2.8 0.28 1.30 1.4
CR/LDUAL + DUL Ty = 0.036H 1–12 2.8 0.17 1.16 1.4
CR/LDUAL + DUM Ty = 0.034H 1–12 2.8 0.21 1.28 1.4
CR/LDUAL + DUH Ty = 0.033H 1–12 2.8 0.26 1.40 1.4
S/LFM + DUM Ty = 0.091H 0.8 1–12 2.8 0.75 3.50 1.33 Abou-Elfath et al. (2017), FEMA (2003, 2014),
S/LFM + DUH Ty = 0.091H 0.8 1–12 2.8 0.90 5.00 1.33 Gupta and Krawinkler (2000), Ramirez et al.
(2012)
S/LFBR + DUM Ty = 0.072H 0.8 1–12 2.8 0.63 2.80 1.33
S/LFBR + DUH Ty = 0.072H 0.8 1–12 2.8 0.75 4.00 1.33
S/LWAL + DUM Ty = 0.053H 0.75 1–12 2.8 0.25 1.45 1.33
S/LWAL + DUH Ty = 0.053H 0.75 1–12 2.8 0.31 2.40 1.33
S/LFINF + DUM Ty = 0.06H 0.75 1–12 2.8 0.38 1.80 1.33
S/LFINF + DUH Ty = 0.06H 0.75 1–12 2.8 0.47 2.80 1.33
SCR/LDUAL + DUM Ty = 0.051H 1–15 2.8 0.95 4.77 1.4 Plumier and Doneux (2001), Liu et al. (2010),
SCR/LDUAL + DUH Ty = 0.051H 1–15 2.8 1.18 5.96 1.4 Tondini et al. (2018)
Bulletin of Earthquake Engineering
Table 1  (continued)
Building class Ty [s] No. of storeys Inter-sto- Yield drift [%] Ult. drift [%] Γ References
rey height
[m]

W/LFM + DUL Ty = 0.123H 0.54 1–6 2.8 0.43 1.19 1.33 Villar-Vega et al. (2017), Vásquez et al.
(2012),FEMA (2014), Camelo (2003)
W/LFM + DUM Ty = 0.123H 0.54 1–6 2.8 0.83 1.78 1.33
W/LFM + DUH Ty = 0.123H 0.54 1–6 2.8 1.23 2.38 1.33
( )
W + WWD/LWAL + DNO Ty = 0.75 ⋅ 0.066H 0.9 1–3 2.6 0.16 1.19 1.4
Bulletin of Earthquake Engineering

( )
+0.25 ⋅ 0.123H 0.54

H height [m], N number of storeys

13
Bulletin of Earthquake Engineering

Fig. 5  Example of capacity curves of bare frame vs infilled frames considering reduction on initial stiffness

Fig. 6  Damage thresholds

damage was assumed to begin at 75% of the yielding displacement (Villar-Vega et al. 2017).
Complete damage was considered to be reached at the ultimate displacement capacity of the
structure. The intermediate damage states (i.e. moderate—DS2 and extensive damage—DS3)
were considered to be evenly spaced between the first and last damage state as depicted in
Fig. 6. This damage criterion follows closely the proposal originally presented by Lagomars-
ino and Giovinazzi (2006). The aleatory uncertainty around the damage criteria was also con-
sidered in the development of the fragility functions. In this process, the engineering demand
parameter for each damage state is assumed to follow a lognormal distribution with a mean
equal to the threshold presented in Fig.  6, and a coefficient of variation. Several existing

13
Bulletin of Earthquake Engineering

Fig. 7  Scatter of ln(EDP) and ln(IM), with the associated best-fit linear curve

studies (e.g. Dymiotis et al. 1999; Borzi et al. 2008a, b; Tarque et al. 2012) have suggested
coefficients of variation for the damage thresholds between approximately 30% and 60%. For
the purpose of this study, a coefficient of variation of 45% was adopted.
Similar to the work of Crowley et al. (2017), the derivation of the fragility functions was
performed following the cloud analysis approach (Jalayer et  al. 2015). This methodology
requires the definition of a best fit curve between an intensity measure (IM) and an engineer-
ing demand parameter (EDP) in the logarithmic space. An example of this statistical regres-
sion is depicted in Fig. 7 for a 2-storey reinforced masonry structure subjected to 300 ground
motion records.
The expected EDP given an IM (E[ln(EDPi)]) and the respective uncertainty due to the
record-to-record variability (σrec_to_rec) can be computed from Eq. (3).

⎧ E[ln EDP]
⎪ �= ln a + b ln IM

(3)
n
⎨ (ln EDPi −E[ln EDP])
2

⎪ 𝜎rec_to_rec = i=1
⎩ n−2

In numerical analyses, it can happen that convergence is attained at levels of deformation


for which the structure is so severely damaged that for all practical purposes a complete loss
must be considered. These high levels of deformation are only reached due to limitations in
the numerical models and numerically instability. Therefore, considering directly these data
points in the estimation of the best fit curve would result in bias in the fragility and vulner-
ability models. For this reason, a censored regression method (Schnedler 2005) was followed
for the calculation of the parameters of the best-fit curve (a and b, as defined in Eq. (3)). These
parameters can be computed using the Maximum Likelihood Method (MLE) (Stafford 2008;
Lallemant et al. 2015), as explained in Eq. (4). This method aims at estimating the parameters
of the best-fit curve (a and b) that maximize the probability of occurrence of the observed data
(i.e. EDPi)

13
Bulletin of Earthquake Engineering

Fig. 8  Scatter of ln(EDP) and ln(IM), with the associated best-fit uncensored and censored regressions

[ ( )]

n
EDPi − a − bIM i
L= 𝜙 (4)
i=1
𝜎

where ϕ stands for the probability density function of the standard Gaussian distribution.
However, as previously mentioned, the structural response in some of the analyses will
be unrealistically high. It is thus convenient to establish a certain threshold (EDPc) above
which the analysis is assumed to have led to collapse or complete damage of the structure.
For the purposes of this study, we have set this threshold to 1.5 times the EDP for complete
damage. At this stage, the observed data can be split between uncensored (­no) and cen-
sored ­(nc) observations. These data points are illustrated in Fig. 8 using circles and squares,
respectively. The uncensored observations are used within the MLE method as defined by
Eq. (4), whist for the censored observations we minimize the negative log-likelihood func-
tion (e.g. Stafford 2008), as defined in Eq. (5).
nc [ ( ( ))] no [ ( )]
∑ EDPc − a − bIM j ∑ EDPi − a − bIM i
ln (L) = ln 1 − 𝛷 + ln 𝜙
j=1
𝜎 i=1
𝜎
(5)
where 𝛷 stands for the cumulative function of the standard Gaussian distribution. In practical
terms, the best-fit curve will be located close to the peak of the probability density function of
the uncensored observations (no) and below the probability density function of the censored
observations (nc). Figure 8 illustrates the data points for the 2-storey reinforced masonry struc-
ture, along with the two best-fit curves: uncensored (considering all observations and Eq. (4))
and censored (splitting the observations between censored/uncensored and using Eq. (5)).
As previously mentioned, the term σrec_to_rec represents the variability due to the record-
to-record variability. In other words, two records with the exact same IM might lead to dis-
tinct structural responses (e.g. Sousa et al. 2018). However, it does not account for another
important source of variability in fragility analysis of building classes: the building-to-
building variability. This source of uncertainty can be introduced in the fragility-derivation
framework by considering a large number of numerical models (e.g. Erberik 2008a, b; Silva

13
Bulletin of Earthquake Engineering

Fig. 9  Construction of a fragility function following the cloud analysis approach

et al. 2014a, b; Villar-Vega et al. 2017) for each building class, or by increasing the standard
deviation directly. The former approach allows considering explicitly the influence of assets
with varying structural and dynamic properties, but it increases considerably the computa-
tional burden. Moreover, the automatic generation of the numerical models also has limita-
tions. The latter approach is simpler, but it requires leveraging upon other studies which
have measured the contribution of this source of variability. For the sake of simplicity, we
have decided to increase directly the standard deviation around the best fit curve by a factor
of 0.30 (e.g. Casotto et al. 2015; Silva et al. 2013; FEMA 2014) as described in Eq. (6):

( )2 ( )2
𝜎total = 𝜎rec_to_rec + 𝜎bld_to_bld (6)

Considering the above, the probability that a given damage state will be reached or
exceeded given an intensity measure (P[DS ≥ dsi|IM]) can be computed as indicated in Eq. (7)
( )
[ ] ln EDPdsi − E[ln EDP| ln IM]
P DS ≥ dsi |IM = 𝛷 (7)
𝜎total

where EDPdsi stands for the engineering demand parameter (i.e. maximum displacement)
for damage state dsi. As previously explained, this threshold is not deterministic, and
instead it follows a lognormal distribution with a mean and coefficient of variation. The
propagation of this source of variability is numerically expressed in Eq. (8)
( )
[ ] ∑ n
ln EDPdsi ,j − E[ln EDP| ln IM] [ ]
P DS ≥ dsi |IM = 𝛷 ⋅ P EDPdsi ,j (8)
j=1
𝜎 total

where EDPdsi,j is an increment of the possible range of EDPs for damage state dsi, and
P[EDPdsi,j] represents the probability of occurrence of EDPdsi,j. The range of possible val-
ues for the EDP and the associated probability of occurrence are estimated using the proba-
bilistic model for each damage state previously described (see Fig. 6). The estimation of
the probability of exceeding moderate damage is illustrated in the left panel of Fig. 9, while
the resulting probability of moderate damage per IM is plotted on the right panel. The
shaded area under the probability density functions illustrated in the left panel represent
the probability of exceeding moderate damage for that particular damage threshold. These
probabilities are represented as data points on the right panel.
Four intensity measures (PGA, SA(0.3 s), SA(0.6 s) and SA(1.0 s)) were considered for the
regression. The minimum distance between the yield period and the period used to estimate

13
Bulletin of Earthquake Engineering

Table 2  Damage to loss model Damage state Expected loss ratio Coefficient


used to derive vulnerability E[LR] of variation
functions
Slight damage (DS1) 0.05 0.30
Moderate damage (DS2) 0.20 0.20
Extensive damage (DS3) 0.60 0.10
Complete damage (DS4) 1.00 0.00

the IM determined the intensity measure to be used in the regression analysis. The fragil-
ity and vulnerability models in this study are intended to be used in large scale seismic risk
assessment studies, and therefore when selecting the intensity measures for the regression
analysis, computational runtime and practicality were prioritized.
In addition to these common IMs, a more robust IM (i.e. average spectral accelera-
tion, AvgSA) (e.g. Eads et al. 2015) was also considered. AvgSA is computed from the geo-
metrical mean of the spectral ordinates in a given interval of periods (usually defined as
0.2T < T<1.5T). The centre period of the interval can be set to any value considered meaning-
ful for the application (e.g. ­T1 or ­Ty). However, this study followed the recommendations from
Kohrangi et al. (2018), and therefore a discrete number of periods (T∊ [0.5, 0.75, 1, 1.5, 1.75,
2, 2.25, 2.5] [s]) was selected to optimize computational runtime. The centre period for each
building class was selected from the list of Kohrangi et al. (2018) by minimizing the distance
to the respective yield period. Once selected the centre period, the average spectral accelera-
tion was computed considering 10 discrete periods evenly spaced between 0.2T and 1.5T
Converting fragility into vulnerability functions is then numerically performed through
Eq. (9) using a damage-to-loss model that correlates each damage state with the respective
probability distribution of loss ratio. The damage-to-loss model was defined based on some
existing models (e.g. Di Pasquale and Goretti 2001; FEMA 2003; Kappos et al. 2006; Bal
et al. 2010a, b), as described in Table 2. The distribution of loss ratio per damage state was
assumed to follow a beta distribution (e.g. Dolce et  al. 2006; Martins et  al. 2016), with
a coefficient of variation defined based on the range of values proposed by Kappos et al.
(2006), the variability in insurance claims described by Bal et al. (2010b), and the distribu-
tion of loss ratios estimated for South American events (Cabrera et al. 2019)

∑ ∑( [
nDS m
] [ ])
E[LR|IM] = P DS = dsi |IM ⋅ LRi,j ⋅ P LRi,j (9)
i=1 j=1

where P[DS = dsi|IM] is the probability of occurrence of damage state dsi given an IM


(computed from the fragility functions), and LRi,j and P[LRi,j] stand for a possible loss ratio
for damage state dsi and associated probability of occurrence (computed from the associ-
ated beta distribution).

4 Fragility and vulnerability results

Due to the large number of building classes analysed in this study, only the results for
unreinforced masonry structures are shown in this section as it is one of the most common
building classes worldwide. Additional information can be found on the public repository3

3
  https​://githu​b.com/lmart​ins88​/globa​l_fragi​lity_vulne​rabil​ity.

13
Bulletin of Earthquake Engineering

Table 3  Fragility parameters for masonry building classes


Building class IM DS1 DS2 DS3 DS4
θ β θ β θ β θ β

MUR/LWAL + DNO/H1 SA(0.3 s) 0.53 0.60 1.05 0.60 1.49 0.60 1.88 0.60
MUR/LWAL + DNO/H2 SA(0.3 s) 0.43 0.58 0.91 0.58 1.36 0.58 1.78 0.58
MUR/LWAL + DNO/H3 SA(0.3 s) 0.41 0.57 0.96 0.57 1.47 0.57 1.95 0.57
MUR/LWAL + DNO/H4 SA(0.6 s) 0.18 0.56 0.46 0.56 0.74 0.56 1.02 0.56
MUR/LWAL + DNO/H5 SA(0.6 s) 0.20 0.54 0.50 0.54 0.80 0.54 1.09 0.54
MUR + ADO/LWAL/DNO/H1 SA(0.3 s) 0.34 0.59 0.73 0.59 1.04 0.59 1.33 0.59
MUR + ADO/LWAL/DNO/H2 SA(0.3 s) 0.27 0.56 0.66 0.56 1.02 0.56 1.35 0.56
MUR + ADO/LWAL/DNO/H3 SA(0.6 s) 0.12 0.54 0.32 0.54 0.51 0.54 0.70 0.54
MUR + STRUB/LWAL + DNO/H1 SA(0.3 s) 0.37 0.60 0.81 0.60 1.16 0.60 1.48 0.60
MUR + STRUB/LWAL + DNO/H2 SA(0.3 s) 0.36 0.56 0.77 0.56 1.15 0.56 1.50 0.56
MUR + STRUB/LWAL + DNO/H3 SA(0.3 s) 0.35 0.58 0.82 0.58 1.24 0.58 1.65 0.58
MUR + STRUB/LWAL + DNO/H4 SA(0.6 s) 0.16 0.54 0.39 0.54 0.61 0.54 0.84 0.54
MUR + STRUB/LWAL + DNO/H5 SA(0.6 s) 0.17 0.53 0.42 0.53 0.68 0.53 0.94 0.53
MUR + STDRE/LWAL + DNO/H1 SA(0.3 s) 0.50 0.60 0.92 0.60 1.29 0.60 1.61 0.60
MUR + STDRE/LWAL + DNO/H2 SA(0.3 s) 0.36 0.58 0.81 0.58 1.21 0.58 1.59 0.58
MUR + STDRE/LWAL + DNO/H3 SA(0.3 s) 0.40 0.58 0.87 0.58 1.31 0.58 1.72 0.58
MUR + STDRE/LWAL + DNO/H4 SA(0.6 s) 0.16 0.56 0.40 0.56 0.64 0.56 0.88 0.56
MUR + STDRE/LWAL + DNO/H5 SA(0.6 s) 0.18 0.53 0.45 0.53 0.72 0.53 0.99 0.53
MUR + CL99/LWAL + DNO/H1 SA(0.3 s) 0.47 0.59 0.87 0.59 1.21 0.59 1.51 0.59
MUR + CL99/LWAL + DNO/H2 SA(0.3 s) 0.35 0.56 0.81 0.56 1.23 0.56 1.63 0.56
MUR + CL99/LWAL + DNO/H3 SA(0.3 s) 0.38 0.59 0.87 0.59 1.32 0.59 1.74 0.59
MUR + CL99/LWAL + DNO/H4 SA(0.6 s) 0.17 0.55 0.42 0.55 0.68 0.55 0.93 0.55
MUR + CL99/LWAL + DNO/H5 SA(0.6 s) 0.18 0.53 0.46 0.53 0.73 0.53 1.01 0.53
MUR + CB99/LWAL + DNO/H1 SA(0.3 s) 0.53 0.60 1.16 0.60 1.70 0.60 2.19 0.60
MUR + CB99/LWAL + DNO/H2 SA(0.3 s) 0.41 0.58 0.98 0.58 1.50 0.58 1.99 0.58
MUR + CB99/LWAL + DNO/H3 SA(0.3 s) 0.48 0.56 1.08 0.56 1.62 0.56 2.15 0.56
MUR + CB99/LWAL + DNO/H4 SA(0.6 s) 0.20 0.56 0.52 0.56 0.83 0.56 1.14 0.56
MUR + CB99/LWAL + DNO/H5 SA(0.6 s) 0.53 0.60 1.16 0.60 1.70 0.60 2.19 0.60

PGA peak ground acceleration, SA(X) spectral acceleration at T = X s, θ median, β logarithmic standard
deviation

regarding all of the statistical parameters per fragility curve. Table 3 provides the fragility
parameters (i.e. median and logarithmic standard deviation) for the unreinforced masonry
classes in the database, while Fig. 10 presents some examples of fragility and vulnerability
plots.

5 Verification and discussion of results

Special attention was given to the validation of the fragility and vulnerability models.
A literature review was performed to assess reasonable values for the dispersion (i.e.
β). As demonstrated by Haselton et al. 2011; Lazar and Dolšek 2014; Silva et al. 2014b

13
Bulletin of Earthquake Engineering

Fig. 10  Example of fragility functions in Sa (top), AvgSa (middle), and vulnerability functions (bottom) for
some masonry classes

for regular buildings the β values for complete damage state should vary between 0.30
and 0.80, with other damage states expected to have slightly lower values. The β values
­(minβ = 0.39, ­maxβ = 0.61) obtained in this study verify that the curves comply with this
preliminary verification criterion.
A second level of verification was performed by computing the expected average annual
loss ratio (AALR) for 13 locations (see Fig.  11) around the world with distinct seismic
hazard levels (San Francisco, Oakland, San Jose, Sacramento, Los Angeles, San Diego,

13
Bulletin of Earthquake Engineering

Fig. 11  Seismic hazard curves

Bogotá, Caracas, Lisbon, Barcelona, Rome, Vienna and Istanbul). These hazard curves
were derived for the four IMs using the mosaic of seismic hazard models of GEM (Pagani
et al. 2019).
This verification was deemed necessary for early identification of excessively vulnerable
(or excessively resistant) building classes. This also allowed verifying the relative vulnera-
bility of each building class and dismiss any obvious inaccuracies (e.g. for the same number
of storeys reinforced masonry having higher annualized losses than unreinforced masonry).
An example of the AALR results for the unreinforced masonry building classes presented
previously are included in Table 4 considering 3 locations (Vienna, Lisbon and Oakland)
with different levels of seismic hazard (i.e. low, moderate and high hazard, respectively).
The results for the remaining cities are available on the public GitHub repository
In addition to the AALRs per building class listed in Table  4, a comparison between
the AALRs aggregated by macro-building classes (i.e. CR/LDUAL, CR/LFINF, CR/LFM,
CR/LWAL, MCF, MR, MUR, S/LFBR, S/LFINF, S/LFM, S/LWAL and W) is provided in
Fig. 12, for the same 3 locations (Vienna, Lisbon and Oakland).
From Fig.  12 one can observe that, unsurprisingly, unreinforced masonry is the most
vulnerable class, whilst braced steel frames present the lowest vulnerability. The annual
expected loss for the macro-building classes are confined between 0.003% (S/LFBR in
Vienna) and 0.40% (MUR in Oakland). To provide context to these results, it is neces-
sary to evaluate seismic risk disaggregated both by building class and by hazard intensity,
though such results or empirical data are relatively scarce in the literature.
Crespi et al. (2019) examined the last seismic events in Central Italy (namely L’Aquila
2009 and Amatrice 2016) and indicated an AALR in the order of 0.40% for old masonry
buildings in the region. The value proposed by the authors is in line with the AALR esti-
mated for Oakland in Fig. 12. Seismic hazard in Central Italy is only moderate-to-high (i.e.
about 0.3 g for the 475-year return period vs 0.5 g for Oakland for the same return period).
An explanation for the lower values presented in Fig. 12 might be due to the existence of
more fragile old unreinforced masonry classes in the region. Figure  12 provides an esti-
mate for the macro class, which includes both more fragile and less fragile masonry build-
ing classes.

13
Bulletin of Earthquake Engineering

Table 4  Average annual loss ratio for unreinforced masonry classes


Building class IM AALR Vienna AALR Lisbon AALR Oakland
[%] [%] [%]

MUR/LWAL + DNO/H1 SA(0.3 s) 0.03 0.10 0.24


MUR/LWAL + DNO/H2 SA(0.3 s) 0.04 0.12 0.30
MUR/LWAL + DNO/H3 SA(0.3 s) 0.03 0.11 0.27
MUR/LWAL + DNO/H4 SA(0.6 s) 0.05 0.16 0.39
MUR/LWAL + DNO/H5 SA(0.6 s) 0.04 0.13 0.32
MUR + ADO/LWAL/DNO/H1 SA(0.3 s) 0.06 0.21 0.50
MUR + ADO/LWAL/DNO/H2 SA(0.3 s) 0.06 0.22 0.54
MUR + ADO/LWAL/DNO/H3 SA(0.6 s) 0.09 0.31 0.75
MUR + STRUB/LWAL + DNO/H1 SA(0.3 s) 0.05 0.17 0.41
MUR + STRUB/LWAL + DNO/H2 SA(0.3 s) 0.05 0.16 0.40
MUR + STRUB/LWAL + DNO/H3 SA(0.3 s) 0.05 0.18 0.44
MUR + STRUB/LWAL + DNO/H4 SA(0.6 s) 0.06 0.21 0.51
MUR + STRUB/LWAL + DNO/H5 SA(0.6 s) 0.06 0.21 0.50
MUR + STDRE/LWAL + DNO/H1 SA(0.3 s) 0.04 0.13 0.31
MUR + STDRE/LWAL + DNO/H2 SA(0.3 s) 0.05 0.16 0.39
MUR + STDRE/LWAL + DNO/H3 SA(0.3 s) 0.04 0.13 0.32
MUR + STDRE/LWAL + DNO/H4 SA(0.6 s) 0.06 0.20 0.50
MUR + STDRE/LWAL + DNO/H5 SA(0.6 s) 0.06 0.21 0.51
MUR + CL99/LWAL + DNO/H1 SA(0.3 s) 0.04 0.14 0.35
MUR + CL99/LWAL + DNO/H2 SA(0.3 s) 0.04 0.15 0.37
MUR + CL99/LWAL + DNO/H3 SA(0.3 s) 0.04 0.14 0.34
MUR + CL99/LWAL + DNO/H4 SA(0.6 s) 0.05 0.18 0.44
MUR + CL99/LWAL + DNO/H5 SA(0.6 s) 0.05 0.18 0.45
MUR + CB99/LWAL + DNO/H1 SA(0.3 s) 0.02 0.08 0.20
MUR + CB99/LWAL + DNO/H2 SA(0.3 s) 0.03 0.11 0.27
MUR + CB99/LWAL + DNO/H3 SA(0.3 s) 0.03 0.09 0.21
MUR + CB99/LWAL + DNO/H4 SA(0.6 s) 0.04 0.13 0.31
MUR + CB99/LWAL + DNO/H5 SA(0.6 s) 0.04 0.13 0.32

Hwang and Lignos (2017) analysed the performance of steel moment frame office
buildings from 4 to 20 storeys with a strong column-weak beam mechanism in high
hazard regions. The authors concluded that the average annual loss is dominated by
repairs to the non-structural acceleration sensitive elements and could vary from 0.38-
0.74% during the building’s life expectancy. The total loss is further disaggregated in
(i) collapse losses, (ii) demolition losses, (iii) structural repair, (iv) drift sensitive non-
structural repair and (v) acceleration sensitive non-structural repair. Considering only
structural and non-structural drift sensitive losses the AALR drops to less than half (at
around 0.15%) of the initial value. This value is still much higher than the one provided
in Fig.  12, probably due to the fragile collapse mechanism considered by Hwang and
Lignos (2017) and due to the fact that this study considered a much more detailed dis-
aggregation of losses than what has been considered herein. Therefore, this should be
considered as an upper bound for the true value for the AALR of steel construction.

13
Bulletin of Earthquake Engineering

0.45

0.4 Vienna
Lisbon
0.35
Oakland
0.3
AALR [%]

0.25

0.2

0.15

0.1

0.05

Fig. 12  Comparison of AALR for different macro building classes

Asprone et  al. (2013) while developing an insurance model for Italy conducted an
event based seismic risk assessment study and provided some estimates for the expected
annual loss for both masonry and reinforced concrete construction. The data provided
by the authors placed the AALR for masonry construction in 0.37% whilst the expected
annual loss for reinforced concrete was 0.18%. The values proposed by Asprone et al.
(2013) are consistent with the findings from Crespi et al. (2019), and considering that
Italy is a region with moderate-to-high seismic hazard, also in agreement with the find-
ings presented herein.
Haselton et al. (2008) developed a detailed methodology to compute losses for rein-
forced concrete buildings from the expected loss of individual elements (e.g. repairing
structural elements, repairing partitions, repainting, clearing debris). The authors con-
cluded that the expected AALR due to earthquakes could amount to 0.6% of the build-
ing’s replacement cost. Hazelton et al. also added that the annual losses are dominated
by (i) repairing partitions, (ii) repairing structural elements and (iii) repainting the inte-
rior in this order of importance.
Analyses performed at a national or regional scale by governmental agencies may
also provide meaningful information to draw conclusions on Fig.  12. For example,
FEMA P-366 (FEMA 2017) has estimated an annualized earthquake loss for Oakland
metropolitan area in the order of $794 million with a respective loss ratio of 0.144%.
In the same document, it is mentioned that the building stock in the state of California
is dominated by timber construction (77%). From this information, one can expect an
annualized loss ratio for timber construction in the order of 0.1% for the city of Oak-
land, in line with the value estimated in Fig. 12 even if slightly above it.

13
Bulletin of Earthquake Engineering

6 Final remarks

Despite the recent efforts in developing fragility and vulnerability models, a literature
search revealed that reliable vulnerability models are still either non-existent or unavail-
able for a significant number of building classes worldwide. For this reason, the Global
Earthquake Model foundation led an endeavour to develop a database of fragility and
vulnerability models for the most frequent building classes worldwide.
The initial step was to collect the building classes of interest. For this end an online
questionnaire was developed and sent to local experts on seismic risk. The output of
the survey allowed to identify the initial 200 building classes in the database. This list
was later expanded to meet the requirements of the Global Risk Map initiative by dis-
cretizing the data in number of storeys and adding new building classes. The final list
of fragility and vulnerability models comprises nearly 500 building classes and freely
accessible through the OpenQuake platform. The fragility models were developed fol-
lowing the Cloud analysis methodology proposed by Jalayer et al. (2015) and nonlinear
time-history analysis on equivalent SDOF oscillators.
The fragility models developed within this study considered ground motion records
from both subduction and active shallow tectonic environments. As stated in Sect. 3.1,
ideally the risk modeller should consider the distinct tectonic environments when
selecting ground motion records, as the seismicity for a particular location might be
driven by ground motion from a specific tectonic environment. We also acknowledge
that the specific characteristic of the ground motion records from each tectonic environ-
ment can influence the resulting fragility functions, specifically when an inefficient IM
is used. For example, ground motion records for active shallow regions tend to be richer
in accelerations for high frequencies, while records for subduction areas tend to have
slightly higher accelerations for long periods of vibration. However, due to the global
scope of this study this distinction was not considered. Instead, we have decided to
release the tools, datasets and models used in this study so other researchers can adjust
the framework to their specific needs, which may include the selection of ground motion
records for specific locations.
A great emphasis was given in the verification and validation of the fragility and vulner-
ability models proposed in this study. As part of the Global Risk Map (Silva et al. 2019) it
was necessary to ensure that the vulnerability models provided verifiable and reliable seis-
mic risk estimates (e.g. average annual loss ratio). To this end, several stages of verifica-
tion were implemented prior to finalize the curves. In an initial verification to disclose any
gross inaccuracies in the models, the average probability of collapse and average annual
loss was computed for a set of locations with different levels of seismic hazard. The risk
metrics provided by the models were also compared with different studies in the literature.
Only after the vulnerability models yield reliable seismic risk estimates the models were
further calibrated using nation and regional level seismic risk assessment. To this end,
event-based seismic risk assessments were conducted for all the nations. The models were
then fine-tuned to ensure that (i) the expected loss ratios between different building classes
follow a sensible ranking (e.g. decrease of AALR with increasing ductility level), and (ii)
the seismic risk at a national level is compatible with previous estimates and historical
data. The result was a list of nearly 500 vulnerability models that are applicable to the most
common building classes worldwide. This study improves on former similar endeavours by
applying the same analytical methodology to develop all functions (meaning the models
are compatible) and to be verified through practical seismic risk assessment studies.

13
Bulletin of Earthquake Engineering

References
Abou-Elfath H, Ramadan M, Meshaly M, Fdiel HA (2017) Seismic performance of steel frames
designed using different allowable story drift limits. Alexandria Eng J 56(2):241–249. https​://doi.
org/10.1016/j.aej.2016.08.028
Acevedo AB, Jaramillo JD, Yepes C, Silva V, Osorio FA, Villar M (2017) Evaluation of the seismic risk
of the unreinforced masonry building stock in Antioquia, Colombia. Nat Hazards 86(1):31–54.
https​://doi.org/10.1007/s1106​9-016-2647-8
Ahmad N, Crowley H, Pinho R, Ali Q (2010) Displacement-based earthquake loss assessment of
masonry buildings in Mansehra City, Pakistan. J Earthq Eng 14(Sup1):1–37. https​://doi.
org/10.1080/13632​46100​36517​94
Alcocer S, Guillermo J, Vazquez A (2004) Response assessment of Mexican confined masonry struc-
tures through shaking table tests. In: 13th World conference on earthquake engineering, Vancou-
ver, Canada
Asprone D, Jalayer F, Simonelli S, Acconcia A, Prota A, Manfredi G (2013) Seismic insurance model for the
Italian residential building stock. Struct Saf 44:70–79. https​://doi.org/10.1016/j.strus​afe.2013.06.001
Bal IE, Bommer JJ, Stafford PJ, Crowley H, Pinho R (2010a) The influence of geographical resolution of
urban exposure data in an earthquake loss model for Istanbul. Earthq Spectra 26(3):619–634. https​://
doi.org/10.1193/1.34591​27
Bal, İ.E., Crowley, H. and Pinho, R. (2010b) Displacement-based earthquake loss assessment: method
development and application to turkish building stock. Research Report No. ROSE-2010-02, Instituto
Universitario di Studi Superiori di Pavia
Bommer J, Spence R, Erdik M, Tabuchi S, Aydinoglu N, Booth E, del Re D, Peterken O (2002) Develop-
ment of an earthquake loss model for Turkish catastrophe insurance. J Seismol 6(3):431–446. https​://
doi.org/10.1023/a:10200​95711​419
Bommer JJ, Stafford PJ, Alarcón JE (2009) Empirical equations for the prediction of the significant, brack-
eted, and uniform duration of earthquake ground motion empirical equations for the prediction of
the duration of earthquake ground motion. Bull Seismol Soc Am 99(6):3217–3233. https​://doi.
org/10.1785/01200​80298​
Borzi B, Crowley H, Pinho R (2008a) Simplified pushover-based earthquake loss assessment (SP-BELA)
method for masonry buildings. Int J Archit Herit 2(4):353–376. https​://doi.org/10.1080/15583​05070​
18281​78
Borzi B, Pinho R, Crowley H (2008b) Simplified pushover-based vulnerability analysis for large-
scale assessment of RC buildings. Eng Struct 30(3):804–820. https​://doi.org/10.1016/j.engst​
ruct.2007.05.021
Brzev S, Scawthorn C, Charleson LA, Greene M, Jaiswal K, Silva V (2013) GEM building taxonomy ver-
sion 2.0. Vol. GEM Technical Report 2013-02 V1.0.0. Pavia, IT: GEM Foundation. 188
Bull DK (2013) Earthquakes and the effects on structures: some of the lessons learnt. Aust J Struct Eng
14(2):145–166. https​://doi.org/10.7158/S12-034.2013.14.2
Burby RJ, French SP, Nelson AC (1998) Plans, code enforcement, and damage reduction: evidence from the
Northridge Earthquake. Earthq Spectra 14(1):59–74. https​://doi.org/10.1193/1.15859​88
Cabrera T, Hube MA, Santa Maria H, Silva V, Martins L, Yepes C, Cortes A (2019) Seismic vulnerability
and fragility curves of houses based on damage data from three earthquakes in Chile. Earthq Spectra,
(in review)
Calvi GM (1999) A displacement-based approach for vulnerability evaluation of classes of buildings. J
Earthq Eng 3(3):411–438
Calvi GM (2010) Engineers understanding of earthquakes demand and structures response. In Geotechnical,
geological and earthquake engineering, p 223–247
Calvi GM, Pinho R (2004) LESSLOSS: a European integrated project on risk mitigation for earthquakes
and landslides. Research report ROSE. University of Pavia Distribuito da IUSS Press, Pavia. vi, 178
Calvi G, Pinho R, Crowley H (2006) State-of-the-knowledge on the period elongation of RC buildings dur-
ing strong ground shaking
Camelo V (2003) Dynamic characteristics of woodframe buildings, Ph.D. thesis, California Institute of
Technology
Casotto C, Silva V, Crowley H, Nascimbene R, Pinho R (2015) Seismic fragility of Italian RC precast indus-
trial structures. Eng Struct 94:122–136. https​://doi.org/10.1016/j.engst​ruct.2015.02.034
Crespi P, Giordano N, Frascaro G (2019) Seismic loss estimation of an old masonry building in Italy. In:
ICASP13—13th international conference on applications of statistics and probability in civil engi-
neering, Seoul, South Korea

13
Bulletin of Earthquake Engineering

Crowley H, Pinho R (2004) Period-Height relationship for existing European reinforced concrete buildings.
J Earthq Eng 8(Sup001):93–119. https​://doi.org/10.1080/13632​46040​93505​22
Crowley H, Pinho R (2006) Simplified equations for estimating the period of vibration of existing buildings.
In: Proceedings of the 1st European conference on earthquake engineering and seismology
Crowley H, Pinho R (2010) Revisiting Eurocode 8 formulae for periods of vibration and their employment
in linear seismic analysis. Earthq Eng Struct Dyn 39(2):223–235. https​://doi.org/10.1002/eqe.949
Crowley H, Pinho R, Bommer JJ (2004) A probabilistic displacement-based vulnerability assessment pro-
cedure for earthquake loss estimation. Bull Earthq Eng 2(2):173–219. https​://doi.org/10.1007/s1051​
8-004-2290-8
Crowley H, Polidoro B, Pinho R, van Elk J (2017) Framework for developing fragility and consequence
models for local personal risk. Earthq Spectra 33(4):1325–1345. https​://doi.org/10.1193/08311​6eqs1​
40m
Crowley H, Rodrigues D, Silva V, Despotaki V, Romao X, Castro JM, Akkar S, Hancılar U, Pitilakis KPD,
Belvaux M, Wiemer S, Danciu L, Correia AA, Bursi OS, Wenzel M (2018) Towards a uniform earth-
quake risk model for Europe. In: 16th European conference on earthquake engineering. Thessaloniki,
Greece
D’Ayala D, Meslem A, Vamvatsikos D, Porter K, Rossetto T (2015) GEM guidelines for analytical vulner-
ability assessment of low/mid-rise buildings
Del Gaudio C, De Risi MT, Ricci P, Verderame GM (2019) Empirical drift-fragility functions and loss esti-
mation for infills in reinforced concrete frames under seismic loading. Bull Earthq Eng 17(3):1285–
1330. https​://doi.org/10.1007/s1051​8-018-0501-y
Di Pasquale G, Goretti A (2001) Vulnerabilità funzionale ed economica degli edifici residenziali colpiti dai
recenti eventi sismici italiani. In: Proceedings of the 10th national conference “L’ingegneria Sismica
in Italia”. Potenza-Matera, Italy
Dolce M, Kappos A, Masi A, Penelis G, Vona M (2006) Vulnerability assessment and earthquake damage
scenarios of the building stock of Potenza (Southern Italy) using Italian and Greek methodologies.
Eng Struct 28(3):357–371. https​://doi.org/10.1016/j.engst​ruct.2005.08.009
Dolšek M (2012) Simplified method for seismic risk assessment of buildings with consideration of alea-
tory and epistemic uncertainty. Struct Infrastruct Eng 8(10):939–953. https​://doi.org/10.1080/15732​
479.2011.57481​3
Dolšek M, Fajfar P (2008) The effect of masonry infills on the seismic response of a four-storey reinforced
concrete frame: a deterministic assessment. Eng Struct 30(7):1991–2001. https​://doi.org/10.1016/j.
engst​ruct.2008.01.001
Dymiotis C, Kappos AJ, Chryssanthopoulos MK (1999) Seismic reliability of RC frames with uncer-
tain drift and member capacity. J Struct Eng 125(9):1038–1047. https​://doi.org/10.1061/
(ASCE)0733-9445(1999)125:9(1038)
Eads L, Miranda E, Lignos DG (2015) Average spectral acceleration as an intensity measure for collapse
risk assessment. Earthq Eng Struct Dyn 44(12):2057–2073. https​://doi.org/10.1002/eqe.2575
Elnashai AS, Di Sarno L (2008) Fundamentals of earthquake engineering. Wiley, Chichester, p 347
Erberik MA (2008a) Fragility-based assessment of typical mid-rise and low-rise RC buildings in Turkey.
Eng Struct 30(5):1360–1374. https​://doi.org/10.1016/j.engst​ruct.2007.07.016
Erberik MA (2008b) Generation of fragility curves for Turkish masonry buildings considering in-plane fail-
ure modes. Earthq Eng Struct Dyn 37(3):387–405. https​://doi.org/10.1002/eqe.760
Erdik M, Aydinoglu N, Fahjan Y, Sesetyan K, Demircioglu M, Siyahi B, Durukal E, Ozbey C, Biro Y,
Akman H, Yuzugullu O (2003) Earthquake risk assessment for Istanbul metropolitan area. Earthq
Eng Eng Vib 2(1):1–23. https​://doi.org/10.1007/BF028​57534​
FEMA (2003) FEMA 450-1 : NEHRP recommended provisions for seismic regulations for new buildings
and other structures, FEMA 450-1, Department of Homeland Security - Federal Emergency Manage-
ment Agency, Washington, DC
FEMA (2014) HAZUS-MH MR5, Technical manual, Department of Homeland Security - Federal Emer-
gency Management Agency
FEMA (2017) FEMA P366-estimated annualized earthquake losses for the United States. FEMA P-366,
Department of Homeland Security - Federal Emergency Management Agency, Washington, DC
Ferreira TM, Maio R, Vicente R, Costa A (2016) Earthquake risk mitigation: the impact of seismic ret-
rofitting strategies on urban resilience. Int J Strategic Prop Manag 20(3):291–304. https​://doi.
org/10.3846/16487​15X.2016.11876​82
García H, Degrande G (2017) Performance and seismic vulnerability of a typical confined masonry house
used in Cuenca Ecuador. In: Proceedings of the 16th World conference on earthquake engineering.
Santiago, Chile

13
Bulletin of Earthquake Engineering

Gautam D, Chaulagain H (2016) Structural performance and associated lessons to be learned from world
earthquakes in Nepal after 25 April 2015 (MW 7.8) Gorkha earthquake. Eng Fail Anal 68:222–243.
https​://doi.org/10.1016/j.engfa​ilana​l.2016.06.002
Ghobarah A (2004) On drift limits with different damage levels. In: Proceedings of international workshop
on performance-based seismic design concepts and implementation. Bled, Slovenia
Grubišić M, Kalman Šipoš T, Sigmund V (2013) Seismic fragility assessment of masonry infilled rein-
forced concrete frames. In: 50th SE-EEE: Skopje, North Macedonia
Gupta A, Krawinkler H (2000) Behavior of ductile SMRFs at various seismic hazard levels. J Struct Eng
126(1):98–107. https​://doi.org/10.1061/(ASCE)0733-9445(2000)126:1(98)
Hamzeh L, Ashour A, Galal K (2018) Development of fragility curves for reinforced-masonry structural
walls with boundary elements. J Perform Constr Facil 32(4):04018034. https​://doi.org/10.1061/
(ASCE)CF.1943-5509.00011​74
Haselton CB, Goulet CA, Mitrani-Reiser J, Beck JL, Dierlein GG, Porter KA, Stewart JP, Taciroglu E
(2008) An assessment to benchmark the seismic performance of a code-confirming reinforced
concrete moment-frame building, PEER report 2007/12, PEER - Pacific Earthquake Engineering
Research Center
Haselton CB, Liel AB, Deierlein GG, Dean BS, Chou JH (2011) Seismic collapse safety of reinforced
concrete buildings. I: assessment of ductile moment frames. J Struct Eng 137(4):481–491. https​://
doi.org/10.1061/(asce)st.1943-541x.00003​18
Hoult R, Goldsworthy H, Lumantarna E (2019) Fragility functions for RC shear wall buildings in Aus-
tralia. Earthq Spectra 35(1):333–360. https​://doi.org/10.1193/12071​7EQS2​51M
Hwang S-H, Lignos DG (2017) Earthquake-induced loss assessment of steel frame buildings with spe-
cial moment frames designed in highly seismic regions. Earthq Eng Struct Dyn 46(13):2141–
2162. https​://doi.org/10.1002/eqe.2898
Jaiswal KS, Wald DJ (2011) Rapid estimation of the economic consequences of global earthquakes.
Open File Report 2011-1116, U.S. Geological Survey
Jaiswal K, Wald D, Porter K (2010) A global building inventory for earthquake loss estimation and risk
management. Earthq Spectra 26(3):731–748. https​://doi.org/10.1193/1.34503​16
Jalayer F, De Risi R, Manfredi G (2015) Bayesian cloud analysis: efficient structural fragility assess-
ment using linear regression. Bull Earthq Eng 13(4):1183–1203. https​://doi.org/10.1007/s1051​
8-014-9692-z
Kappos A, Panagopoulos G, Panagiotopoulos C, Penelis G (2006) A hybrid method for the vulnerability
assessment of R/C and URM buildings. Bull Earthq Eng 4(4):391–413. https​://doi.org/10.1007/
s1051​8-006-9023-0
Karanikoloudis G, Lourenço PB (2018) Structural assessment and seismic vulnerability of earthen his-
toric structures: application of sophisticated numerical and simple analytical models. Eng Struct
160:488–509. https​://doi.org/10.1016/j.engst​r uct.2017.12.023
Kohrangi M, Bazzurro P, Vamvatsikos D, Spillatura A (2017) Conditional spectrum-based ground
motion record selection using average spectral acceleration. Earthq Eng Struct Dyn 46(10):1667–
1685. https​://doi.org/10.1002/eqe.2876
Kohrangi M, Kotha SR, Bazzurro P (2018) Ground-motion models for average spectral acceleration in a
period range: direct and indirect methods. Bull Earthq Eng 16(1):45–65. https​://doi.org/10.1007/
s1051​8-017-0216-5
Lagomarsino S, Giovinazzi S (2006) Macroseismic and mechanical models for the vulnerability and
damage assessment of current buildings. Bull Earthq Eng 4(4):415–443. https​://doi.org/10.1007/
s1051​8-006-9024-z
Lallemant D, Kiremidjian A, Burton H (2015) Statistical procedures for developing earthquake damage
fragility curves. Earthq Eng Struct Dyn 44(9):1373–1389. https​://doi.org/10.1002/eqe.2522
Lazar N, Dolšek M (2014) Incorporating intensity bounds for assessing the seismic safety of structures:
does it matter? Earthq Eng Struct Dyn 43(5):717–738. https​://doi.org/10.1002/eqe.2368
Liel AB, Luco N, Raghunandan M, Champion CP (2015) Modifications to risk-targeted seismic design
maps for subduction and near-fault hazards. In: 12th International conference on applications of
statistics and probability in Civil Engineering, ICASP 2015
Liu J, Liu Y, Liu H (2010) Seismic fragility analysis of composite frame structure based on perfor-
mance. Earthq Sci 23(1):45–52. https​://doi.org/10.1007/s1158​9-009-0049-7
Lotfy I, Mohammadalizadeh T, Ahmadi F, Soroushian S (2019) Fragility functions for displace-
ment-based seismic design of reinforced masonry wall structures. J Earthq Eng. https​://doi.
org/10.1080/13632​469.2019.16598​81

13
Bulletin of Earthquake Engineering

Lovon H, Tarque N, Silva V, Yepes-Estrada C (2018) Development of fragility curves for confined
masonry buildings in Lima, Peru. Earthq Spectra 34(3):1339–1361. https​://doi.org/10.1193/09051​
7eqs1​74m
Martins L, Silva V, Marques M, Crowley H, Delgado R (2016) Development and assessment of damage-
to-loss models for moment-frame reinforced concrete buildings. Earth Eng Struct Dyn 45(5):797–
817. https​://doi.org/10.1002/eqe.2687
Martins L, Silva V, Bazzurro P, Marques M (2018) Advances in the derivation of fragility functions for
the development of risk-targeted hazard maps. Eng Struct 173:669–680. https​://doi.org/10.1016/j.
engst​r uct.2018.07.028
McKenna F, Fenves G, Scott M, Jeremic B (2000) Open system for Earthquake Engineering Simulation
(OpenSees). Pacific Earthquake Engineering Research Center. University of California, Berkeley,
CA
Motamed H, Calderon A, Silva V, Costa C (2018) Development of a probabilistic earthquake loss model
for Iran
Mouroux P, Brun BL (2006) Presentation of RISK-UE project. Bull Earthq Eng 4(4):323–339. https​://
doi.org/10.1007/s1051​8-006-9020-3
Murcia-Delso J, Shing PB (2012) Fragility analysis of reinforced masonry shear walls. Earthq Spectra
28(4):1523–1547. https​://doi.org/10.1193/1.40000​75
Pagani M, Garcia-Pelaez J, Gee R, Johnson K, Silva V, Simionato M, Styron R, Vigano D, Danciu L,
Monelli D, Poggi V, Weatherill G (2019) The 2018 version of the global earthquake model: hazard
component. Earthquake Spectra
Plumier A, Doneux C (2001) Seismic behaviour and design of composite steel concrete structures.
ICONS Report no. 4, LNEC-Laboratorio Nacional de Engenharia Civil, Lisbon, Portugal
Preciado A, Ramirez-Gaytan A, Santos JC, Rodriguez O (2020) Seismic vulnerability assessment and
reduction at a territorial scale on masonry and adobe housing by rapid vulnerability indicators: the
case of Tlajomulco, Mexico. Int J Disaster Risk Reduct 44:101425. https​://doi.org/10.1016/j.ijdrr​
.2019.10142​5
Ramirez CM, Lignos DG, Miranda E, Kolios D (2012) Fragility functions for pre-Northridge welded
steel moment-resisting beam-to-column connections. Eng Struct 45:574–584. https​://doi.
org/10.1016/j.engst​r uct.2012.07.007
Riahi Z, Elwood Kenneth J, Alcocer Sergio M (2009) Backbone model for confined masonry walls for
performance-based seismic design. J Struct Eng 135(6):644–654. https​://doi.org/10.1061/(ASCE)
ST.1943-541X.00000​12
Rossetto T, Elnashai A (2003) Derivation of vulnerability functions for European-type RC structures based
on observational data. Eng Struct 25(10):1241–1263. https​://doi.org/10.1016/s0141​-0296(03)00060​-9
Rossetto T, Ioannou I, Grant D, Maqsood T (2014) Guidelines for empirical vulnerability assessment.
GEM Technical Report 2014-08 V1.0.0, Global Earthquake Model Foundation, Pavia, Italy
Salgado-Gálvez, M., Zuloaga Romero, D., Bernal, G., Mora, M. and Cardona, O. (2013) Probabilistic
seismic risk assessment of the building stock in Medellín, Colombia. In: International symposium,
computational civil engineering. Iasi, Romania
Schnedler W (2005) Likelihood estimation for censored random vectors. Econom Rev 24(2):195–217.
https​://doi.org/10.1081/ETC-20006​7925
Sen TK (2009) Fundamentals of seismic loading on structures. Wiley-Blackwell, Chichester, p 384
Silva V, Crowley H, Pinho R, Varum H (2013) Extending displacement-based earthquake loss assess-
ment (DBELA) for the computation of fragility curves. Eng Struct 56:343–356. https​://doi.
org/10.1016/j.engst​r uct.2013.04.023
Silva V, Crowley H, Varum H, Pinho R (2014a) Seismic risk assessment for mainland Portugal. Bull
Earthq Eng, 1–29. https​://doi.org/10.1007/s1051​8-014-9630-0
Silva V, Crowley H, Varum H, Pinho R, Sousa L (2014b) Investigation of the characteristics of Portu-
guese regular moment-frame RC buildings and development of a vulnerability model. Bull Earthq
Eng, pp 1–36. https​://doi.org/10.1007/s1051​8-014-9669-y
Silva V, Casotto C, Rao A, Villar M, Crowley H, Vamvatsikos D (2015) OpenQuake Risk Modeller’s
toolkit - user guide, Technical Report 2015–09, Global Earthquake Model Foundation, Pavia, Italy
Silva V, Yepes-Estrada C, Dabbeek J, Martins L (2017) GED4ALL—global exposure database for multi-
hazard risk analysis—inception report. GEM Technical Report 2017-01, GEM Foundation, Pavia,
Italy
Silva V, Amo-Oduro D, Calderon A, Costa C, Dabbeek J, Despotaki V, Martins L, Pagani M, Rao
A, Simionato M, Viganò D, Yepes-Estrada C, Acevedo A, Crowley H, Horspool N, Jaiswal K,
Journeay M, Pittore M (2019) Development of a global seismic risk model. Earthq Spectra. https​
://doi.org/10.1177/87552​93019​89995​3

13
Bulletin of Earthquake Engineering

Snoj J, Dolšek M (2017) Fragility functions for unreinforced masonry walls made from hollow clay
units. Eng Struct 145:293–304. https​://doi.org/10.1016/j.engst​r uct.2017.05.001
So E (2016) Estimating fatality rates for earthquake loss models. Springer, New York
Sousa L, Silva V, Marques M, Crowley H (2016) On the treatment of uncertainties in the development
of fragility functions for earthquake loss estimation of building portfolios. Earthq Eng Struct Dyn.
https​://doi.org/10.1002/eqe.2734
Sousa L, Silva V, Marques M, Crowley H (2018) On the treatment of uncertainty in seismic vulnerability
and portfolio risk assessment. Earthq Eng Struct Dyn 47(1):87–104. https​://doi.org/10.1002/eqe.2940
Stafford PJ (2008) Conditional prediction of absolute durationsshort note. Bull Seismol Soc Am 98(3):1588–
1594. https​://doi.org/10.1785/01200​70207​
Strasser FO, Bommer JJ, Şeşetyan K, Erdik M, Çağnan Z, Irizarry J, Goula X, Lucantoni A, Sabetta F, Bal
IE, Crowley H, Lindholm C (2008) A comparative study of european earthquake loss estimation tools
for a scenario in Istanbul. J Earthq Eng 12(sup2):246–256. https​://doi.org/10.1080/13632​46080​20141​
88
Tarque N, Crowley H, Pinho R, Varum H (2010) Seismic risk assessment of adobe dwellings in Cusco,
Peru, based on mechanical procedures. In: 14th European conference on earthquake engineering
Ohrid, North Macedonia
Tarque N, Crowley H, Pinho R, Varum H (2012) Displacement-based fragility curves for seismic assessment
of adobe buildings in Cusco, Peru. Earthq Spectra 28(2):759–794. https​://doi.org/10.1193/1.40000​01
Tomaževič M, Gams M (2012) Shaking table study and modelling of seismic behaviour of confined AAC
masonry buildings. Bull Earthq Eng 10(3):863–893. https​://doi.org/10.1007/s1051​8-011-9331-x
Tomaževič M, Klemenc I (1997) Verification of seismic resistance of confined masonry buildings. Earthq
Eng Struct Dyn 26(10):1073–1088. https​://doi.org/10.1002/(sici)1096-9845(19971​0)26:10%3c107​
3:Aid-eqe69​5%3e3.0.Co;2-z
Tondini N, Zanon G, Pucinotti R, Di Filippo R, Bursi OS (2018) Seismic performance and fragility func-
tions of a 3D steel-concrete composite structure made of high-strength steel. Eng Struct 174:373–383.
https​://doi.org/10.1016/j.engst​ruct.2018.07.026
Ulrich T, Negulescu C, Douglas J (2014) Fragility curves for risk-targeted seismic design maps. Bull Earthq
Eng 12(4):1479–1491. https​://doi.org/10.1007/s1051​8-013-9572-y
Vamvatsikos D, Allin Cornell C (2006) Direct estimation of the seismic demand and capacity of oscillators
with multi-linear static pushovers through IDA. Earthq Eng Struct Dyn 35(9):1097–1117. https​://doi.
org/10.1002/eqe.573
Vásquez L, Hernández G, Campos R, González M (2012) Caracterización mecánica de muros estructurales
de madera, Technical Report No. 191, Instituto Forestal, Chile
Villar-Vega M, Silva V, Crowley H, Yepes C, Tarque N, Acevedo AB, Hube MA, Gustavo CD, María HS
(2017) Development of a fragility model for the residential building stock in South America. Earthq
Spectra 33(2):581–604. https​://doi.org/10.1193/01071​6eqs0​05m
Watson-Lamprey J, Abrahamson N (2006) Selection of ground motion time series and limits on scaling.
Soil Dyn Earthq Eng 26(5):477–482. https​://doi.org/10.1016/j.soild​yn.2005.07.001
Yañez F, Astroza M, Holmerg A, Ogaz O (2004) Behavior of confined masonry shear walls with large open-
ings. In: Proceedings of the 13th World conference on earthquake engineering. Vancouver, Canada
Yepes-Estrada C, Silva V, Rossetto T, D’Ayala D, Ioannou I, Meslem A, Crowley H (2016) The global
earthquake model physical vulnerability database. Earthq Spectra 32(4):2567–2585. https​://doi.
org/10.1193/01181​6eqs0​15dp

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

13

You might also like