You are on page 1of 84

Seismic fragility curves for the

European building stock

Review and evaluation of


analytical fragility curves

Rui Maio, Georgios Tsionis

2015

EUR 27635 EN
Seismic fragility curves for the
European building stock
This publication is a Technical report by the Joint Research Centre, the European Commission’s in-house science
service. It aims to provide evidence-based scientific support to the European policy-making process. The scientific
output expressed does not imply a policy position of the European Commission. Neither the European
Commission nor any person acting on behalf of the Commission is responsible for the use that might be made of
this publication.

Contact information
Name: Georgios Tsionis
Address: Joint Research Centre, Via E. Fermi 2749, Ispra (VA) 21027, Italy
E-mail: georgios.tsionis@jrc.ec.europa.eu
Tel.: 39 03 32 78 94 84

JRC Science Hub


https://ec.europa.eu/jrc

JRC99561

EUR 27635 EN

ISBN 978-92-79-54136-0 (PDF)


ISBN 978-92-79-54137-7 (print)

ISSN 1831-9424 (online)


ISSN 1018-5593 (print)

doi:10.2788/586263 (online)
doi:10.2788/798152 (print)

© European Union, 2015

Reproduction is authorised provided the source is acknowledged.

How to cite: Maio R and Tsionis G (2016) Seismic fragility curves for the European building stock: review and
evaluation of analytical fragility curves. EUR 27635 EN. doi:10.2788/586263.

i
ii
Table of contents

1. Introduction .................................................................................................... 1
1.1. Background ............................................................................................... 1
1.2. Objectives and motivation ........................................................................... 2
1.3. Outline of the report ................................................................................... 3
2. Overview of existing fragility curves ................................................................... 5
2.1. Analytical approaches ................................................................................. 5
2.2. Empirical approaches .................................................................................. 6
2.3. Expert judgment elicitation approaches ......................................................... 7
2.4. Hybrid approaches ...................................................................................... 8
3. Review of analytical fragility assessment methodologies ..................................... 11
3.1. Structural system ..................................................................................... 14
3.2. Components for response analysis .............................................................. 15
3.3. The selection of the analysis type ............................................................... 17
3.3.1. Nonlinear Dynamic Analysis .................................................................. 17
3.3.2. Nonlinear Static Analysis ...................................................................... 19
3.3.3. Simplified Mechanism Models ............................................................... 20
3.3.3.1. FaMIVE method ............................................................................. 21
3.3.3.2. DBELA method .............................................................................. 22
3.3.3.3. SP-BELA method ........................................................................... 23
3.3.3.4. VULNUS method ............................................................................ 25
3.4. The selection of the model type .................................................................. 25
3.4.1. MDoF models ...................................................................................... 26
3.4.2. Reduced MDoF models ......................................................................... 27
3.4.3. SDoF models ...................................................................................... 27
3.5. Shear failure ............................................................................................ 28
3.6. Out-of-plane response of masonry buildings ................................................ 29
3.7. Horizontal diaphragms in masonry buildings ................................................ 30
3.8. Geometrical irregularities........................................................................... 30
3.9. Seismic demand and site-specific records .................................................... 31
3.10. Definition of EDP values at damage thresholds ............................................. 31
3.11. Form of relation ....................................................................................... 32
3.12. Intensity Measure ..................................................................................... 33
3.13. Sampling method and sample size .............................................................. 33
3.14. Summary table of the reviewed fragility curves ............................................ 35
4. Criteria and evaluation of fragility curves .......................................................... 41
4.1. Capacity .................................................................................................. 42

iii
4.2. Demand .................................................................................................. 43
4.3. Fragility curve (derivation method) ............................................................. 43
4.4. Treatment of uncertainty ........................................................................... 44
5. Current trends in seismic fragility curves ........................................................... 49
6. Concluding remarks ........................................................................................ 55

iv
Acknowledgements

The work herein presented entitled Seismic fragility curves for the European building
stock: review and evaluation of analytical fragility curves, was carried out under the
European Commission traineeship job contract no. 228449 at the European Laboratory
for Structural Assessment of the Joint Research Centre, in the framework of the of the
RESURBAN institutional project, which deals with the resilience of the buildings in urban
areas across the European Union. This was not only a quite challenging but also a very
motivating research topic since seismic fragility assessment has becoming an issue of
growing interest worldwide. However, without the support from many sides it would not
have been possible to accomplish this outcome and therefore I would like to use this
occasion to express my gratitude.
First of all I would like to thank my traineeship advisor, Georgios Tsionis, for his support
and understanding, and also for his remarkable comments and suggestions throughout
the traineeship, which were definitely decisive to improve the overall quality of this
report. I would like also to thank Luísa Sousa and Artur Pinto for their support and
willingness to help solving whatever situation, either work-related or everyday life
unexpected difficulties, and Professor Serena Cattari, for her interest, expertise and kind
collaboration on writing the foreword of the present report.
I would like to address a special word of thanks to the Board of Studies of my PhD
Program Infrarisk- “Analysis and Mitigation of risks in Infrastructures”, in particular to
Professor Carlos Sousa Oliveira, for his support and receptiveness regarding this
traineeship opportunity at the Joint Research Centre.
I would also like to thank my PhD supervisor, Professor Romeu Vicente, for his
motivation and positivism that keeps me interested following a research career, and also
for his help and support on overcoming all the institutional and bureaucratic hurdles.
Last but not least, I would like to thank all the extraordinary people that I had the
chance to meet here at the Joint Research Centre, namely to Traineeland. Within this
unofficial community of trainees I have met the most highly talented young individuals
who taught me among other things, to be more tolerant, open-minded and to always
think on the positive side of things.
Rui Maio

PhD Student of the Infrarisk- Doctoral Programme


DECivil-UA, Department of Civil Engineering
RISCO – Aveiro Research Centre of Risks and Sustainability in Construction
University of Aveiro

v
vi
Foreword

Seismic risk is defined as the potential of negative consequences of hazardous events


that may occur in a specific area and period of time and is obtained by the probabilistic
convolution of three main components, the hazard, vulnerability and exposure. The
concept of negative consequences can belong to different categories such as: physical
damage to buildings, casualties, direct cost of damage, indirect economic impacts (loss
of productivity and business interruption), loss of function in lifelines and critical facilities
and also social, organizational and institutional impacts. In recent years a growing
attention in research has been paid on how to consider all these contributions in an
integrate approach within a seismic loss assessment: this also agrees with the emerging
concept of resilience-based assessment. However, vulnerability still represents a
relevant and crucial component of the chain of the seismic risk analysis.
In this context, the so-called fragility curves provide the essential link between the
seismic hazard at the site and the corresponding effects on any kind of exposed
components and aim to relate a selected intensity measure (IM) to a proper engineering
demand parameter.
Many fragility curves are available in literature for ordinary buildings as derived from
different approaches (i.e. empirical, expert elicitation based, analytical, hybrid), but to
select “the most reliable one for a given application” still represents a very challenging
task and the use of existing fragility functions has to be made very carefully.
First of all, the complexity arises from the wide variety of existing constructions, which
are characterized by very different structural systems, moving through historical periods
and geographical areas. This is particularly true for ancient masonry buildings, since
they were often built by an empirical approach that was strongly affected by the
seismicity of the area and by the constructive details specifically conceived to withstand
the earthquake after the damage observation, but it is also valid for the reinforced
concrete ones, which although more engineered, can be influenced by the code rules in
force at the moment of their design or by specific patents.
From that it follows as the first step of the vulnerability assessment at large scale and of
the most reliable selection of fragility curves consists in the proper classification of the
building stock that cannot ignore a preliminary study of the characteristics of the built
environment under investigation. Depending on the availability and reliability of fragility
curves, the building classification should be more or less detailed. In fact, an excessive
splitting of the built environment into detailed classes, with associated low dispersed
fragility curves, turns out to be specious if their reliability is not robust; in these cases it
is better to reduce the number of buildings classes and ascribe to each one a more
reliable fragility curve, even if defined by a bigger dispersion.
As far as the different approaches available to define the fragility curves concern, those
obtained from observed damage after the occurrence of an earthquake (empirical
approach) are surely valuable, because they are directly correlated to the actual seismic
behaviour of buildings and can be very useful for validation of analytical methods and
calibration of hybrid fragility curves. However, empirical fragility curves are strongly
influenced by the reliability of the damage assessment, which is often made by a quick
survey aimed to other scopes, as the building tagging for use and occupancy. Moreover,
they usually refer to the macroseismic intensity, which is not an instrumental measure
but is based on a subjective evaluation. Then, for more accurate loss estimation, it is
often necessary to convert macroseismic intensity into an instrumental intensity
measure, introducing important approximation and normally huge uncertainties. Finally,
it is important to take into account they are obviously affected also by the “local seismic
culture” of the area and by the specific constructive details which characterize the
building stock, thus their extrapolation to other geographic areas have to be made
carefully.

vii
The use of the analytical fragility curves looks quite attractive in moving towards a more
versatile approach since presents the following main advantages: a) fully employ all
results of Probabilistic Seismic Hazard Analysis (instrumental IMs, seismic input in the
spectral form); b) explicitly keep into account the various parameters that determine the
structural response; c) accurately evaluate the uncertainty propagation. Within this
context, fragility curves may be based on nonlinear static approaches through simplified
and detailed models or based on linear dynamic approaches. Although the reliability of
the vulnerability assessment is affected by the capability of the model to simulate the
actual seismic response of the examined class, the use of the nonlinear static approach
through simplified mechanical based models certainly aims to find a compromise among
the different needs to limit the computational effort, to limit the number of parameters
to be acquired for the building stock at large scale, and provide an adequate versatility
to capture different structural responses.
Therefore, within such context and focusing in particular on the analytical approach, this
report aims to contribute in providing useful and effective tools for supporting the
selection of fragility curves available in literature.

Serena Cattari

Assistant Professor
DICCA - Department of Civil, Chemical and Environmental Engineering
University of Genova

viii
Abstract

Earthquakes continue to represent a serious threat for some European countries,


particularly for Mediterranean bordering countries, where these events have been
triggering significant destruction and loss over the last decades. Despite the
unpredictable nature of earthquakes, seismic risk assessment should be addressed
having in mind the main cause of damage, which is related to the inadequate seismic
resistance of the existing structures such as residential, industrial or cultural heritage
buildings.
A literature review on the existing methodologies for deriving fragility curves suitable to
the European building stock is herein presented. Even though a brief overview of all the
existing approaches for deriving fragility curves is made, this report focuses on analytical
fragility curves and discusses the most relevant features inherently associated with the
computation of these curves. Additionally, these methodologies are qualitatively
evaluated by means of different sets of criteria.
With this report it is intended to provide a clear insight about the main differences
between existing analytical methodologies for deriving fragility curves, highlighting as
well some of their most important advantages and drawbacks. The proposed evaluation
criteria can be used further on to help not only on the selection of the most suitable
fragility curves for a given geographical location and structural typology, but also for the
further comparison and validation of fragility curves.

ix
x
1. Introduction

1.1. Background
Earthquakes continue to represent a serious threat for some European countries,
particularly for Mediterranean bordering countries, where these events have been
triggering significant destruction and loss over the last decades. Despite the
unpredictable nature of earthquakes, seismic risk assessment should be addressed
having in mind the main cause of damage, which is related to the inadequate seismic
resistance of the existing structures such as residential, industrial or cultural heritage
buildings.
The causes and consequences associated to seismic risk have been acknowledged by the
European Union, which has expressed great concern about this issue, either by
supporting the development and implementation of Eurocodes, supplementary
coordination of Civil Protection bodies, or even funding numerous research programmes
in this particular field. However, there is still much to be done as regards seismic risk
assessment and mitigation, particularly regarding the oldest and most vulnerable
buildings erected without anti-seismic provisions.
In seismic risk assessment, vulnerability curves are commonly used to express the
likelihood that assets at risk will sustain different degrees of loss over a range of
earthquake ground motion intensities. These vulnerability curves are based on the
statistical analysis of loss values recorded in past earthquakes, simulated in analytical or
numerical methodologies, assumed through expert judgement elicitations, or on a
combination of these methodologies. They can be constructed either by direct or indirect
procedures. While direct procedures usually correlate loss (cost of the physical damage
to buildings, associated casualties and downtime) with a measure of the ground motion
intensity, indirect procedures are usually two-step techniques. In the first step, the
fragility assessment is carried out correlating the physical damage with a measure of the
ground motion intensity. Then, in a second step, fragility curves are transformed in
vulnerability curves through appropriate damage-to-loss functions (D’Ayala et al. 2014).
With very limited exceptions, analytical vulnerability curves make use of indirect
procedures, as they are based on numerical modelling of differing complexity to simulate
the physical phenomenon of seismic damage of the type of structure under analysis. In
Rossetto et al. (2013) and D’Ayala et al. (2014), a discussion concerning the conversion
from physical damage to economic loss and the choice of appropriate damage to loss
functions in relation to the different vulnerability models adopted is made.
It is within this context that fragility curves are becoming more and more important as a
valuable tool for the seismic risk assessment, establishing the link between the seismic
hazard at a site and the effects of the predicted ground motions on the built
environment. By definition, fragility curves express continuous relationships between the
probabilities associated to an asset or class of assets, of reaching or exceeding
predefined damage states for a range of earthquake ground motion intensities.
As reported in literature, fragility curves can be estimated through empirical, analytical,
expert judgment elicitation or hybrid approaches (see Chapter 2). In this report, only
analytically-derived fragility curves were considered, giving particular emphasis on the
variety of methodologies used for their construction, and the corresponding implications
on robustness and reliability aspects that lead to a more systematic qualitative
assessment. According to Calvi et al. (2006), analytical methodologies produce slightly
more detailed and transparent assessment algorithms with direct physical meaning, that
not only allow detailed sensitivity studies to be undertaken, but also cater for the
straightforward calibration to various characteristics of the building stock and hazard. As
one can perceive by the plentiful references made in the present report, several studies

1
have been carried out over the last decades in multiple countries in the field of fragility
assessment for different building typologies.
However, the employment of these existing models in seismic risk analysis may
represent a challenging task due to the fact that the different approaches and
methodologies are spread across scientific journals, conference proceedings, technical
reports and software manuals, hindering the creation of an integrated framework that
could allow the visualization, acquisition and comparison between all the existing curves.
Another important issue stressed out by many researchers is related to the way in which
these curves are defined and presented. The statistical parameters employed to
characterize each curve can vary considerably among different approaches and
methodologies, and in some cases only plots of the final results are provided, instead of
the actual numerical values and respective calculation processes. Finally, it is also worth
referring that it is not often clear what were the methodologies and assumptions
followed in the derivation of the curves, which hinders the adequate evaluation of the
reliability, accuracy and overall quality of the resulting model.
As a response to these concerns, the Global Earthquake Model (GEM) Foundation has
supported the development of an online platform to store, visualize and explore a
multitude of models required to characterize the physical vulnerability of assets. In
addition to the fragility and vulnerability curves, it is also possible to explore damage-to-
loss models and capacity curves. The development of this database relied strongly on
the outcomes of the Global Vulnerability Consortium (GVC) project (Porter et al. 2012)
launched by GEM, which includes the guidelines for developing analytical (D’Ayala et al.
2014; Porter et al. 2014) and empirical (Rossetto et al. 2014) fragility and vulnerability
curves, as well as recommendations for selecting existing empirical and analytical
fragility curves (D’Ayala and Meslem 2012; Rossetto et al. 2013). The fragility models
collected within the European collaborative research project SYNER-G funded by
European Commission (Pitilakis et al. 2014) were also considered for the creation of this
database, which together with the respective recommendations were the starting point
for developing this report that reviews additional methodologies for deriving fragility
curves and applies the selection criteria with a number of modifications.

1.2. Objectives and motivation


Keeping this background in mind, the current report aims to provide additional insight
regarding the following central question: how can one select suitable (sets of) fragility
curves for a given building typology and location from the literature?
The Joint Research Centre (JRC) by means of the European Laboratory for Structural
Assessment (ELSA) has been collaborating and carrying out several research projects
addressing this issue, as for instance the aforementioned SYNER-G project (Pitilakis et
al. 2013). In that project, an integrated methodology for the systemic seismic
vulnerability and risk analysis of buildings, transportation and utility networks and
critical facilities, considering the interactions between different components and systems,
was developed, together with the development of inventory datasets through remote
sensing and direct observation data for earthquake loss estimation, guidelines for
typology definition of European physical assets for seismic risk assessment, among other
issues (Franchin 2013; Hancilar and Taucer 2013; Kaynia 2013; Pitilakis et al. 2013).
Moreover, acknowledging the sustainable development as one of the fundamental
objectives of the European Union, a research project concerning the seismic performance
assessment addressing sustainability and energy efficiency was carried out within the
European Laboratory for Structural Assessment (Tsimplokoukou et al. 2014).
Furthermore, the seismic strengthening of RC buildings and the proposal of new
European technical rules for the assessment and retrofitting of existing structures from
support to the implementation, harmonization and further development of the

2
Eurocodes, have been also addressed in recent years, through the work developed by
Tsionis et al. (2014) and Luechinger and Fischer (2015), respectively.
The work presented herein was carried out in the framework of the RESURBAN
institutional project, which deals with the resilience of the buildings in urban areas
across the European Union (EU), lending continuity to the research developed in recent
years. It relates to the European strategy for disaster management and to the European
Urban Agenda. The work builds on the experience acquired in the field of vulnerability
assessment of buildings through previous experimental and numerical research at the
JRC and aims to provide scientific support for decision-making as regards the seismic
retrofit of existing buildings in the EU and the consequent reduction of losses due to
natural hazards. The report is focussed on the vulnerability of exposed assets, which
constitutes one of the three main components of seismic risk assessment. The fragility
curves reviewed in this work will be used in a probabilistic risk assessment method,
together with the most up-to-date seismic hazard map and an appropriate databank of
the building stock. Different retrofit scenarios will be examined at a later stage,
considering synergies and conflicts with energy retrofitting of buildings.
Aiming to leverage upon the wealth of existing approaches and methodologies that have
been developed in the last decades by dozens of scientists and practitioners, this report
provides a clear insight on the existing literature concerning the derivation of analytical
fragility curves applicable to the European building stock. Furthermore, this report aims
not only to aid the selection among existing fragility curves available for a given
geographic location, structural typology and level of expected hazard, but also to
contribute for the comparison and validation of new fragility curves proposals.

1.3. Outline of the report


This introductory Chapter 1 presents the background and motivations that served as
basis for this report, pointing out the issues to be addressed, highlighting the main
opportunities associated to its development and summarising the most relevant and
recently developed methodologies. Moreover, the general and specific objectives of the
RESURBAN research project are discussed.
In Chapter 2, an overall review of existing fragility curves is carried out, covering the
most common approaches usually employed for the generation of fragility curves
worldwide: empirical; analytical; expert judgment elicitation, and hybrid approaches.
Chapter 3 provides a comprehensive review of the analytical methodologies for fragility
assessment that are available in literature and are applicable to the European and
Mediterranean building stock.
In turn, in Chapter 4 it is proposed a qualitative evaluation criteria based on four main
categories. These criteria are then applied to all fragility curve methodologies herein
reviewed.
In Chapter 5, the main trends of existing analytical methodologies for deriving fragility
curves are identified through statistical analysis of the main factors influencing the
reviewed methodologies and the respective outputs from the qualitative evaluation
carried out in the previous Chapter 4.
Finally, Chapter 6 presents a general overview of the work developed throughout this
traineeship and summarises the key conclusions that have been pointed out in the
previous chapters of the report. To conclude, possible future research lines are also
outlined and discussed.

3
4
2. Overview of existing fragility curves

In this Chapter, an overall review of existing fragility curves will be carried out, aiming to
cover all of the approaches usually employed for the generation of fragility curves
worldwide.
Although several pioneering methodologies for the construction of fragility curves have
been developed worldwide, namely in Japan and in the United States of America, special
care should be given when assigning default curves to represent the seismic
performance of the structure under study or the variability of features determining the
corresponding fragility curves. For instance, the existing HAZUS capacity curves (FEMA
2003a) derived for buildings in the USA, have been widely used to generate fragility
curves of buildings located in different regions of the world, which are designed and built
according to very different standards and construction techniques. Some of these cases
can be found in Lourdes et al. (2007) or Vacareanu et al. (2007), for example. This is
particularly recurrent when studies are conducted for large portions of the building stock
and resources for direct survey and data acquisition are modest. In fact, differences in
construction techniques and detailing between different countries can be significant,
even when buildings are nominally designed to the same code standards (D’Ayala and
Meslem 2013). Such detailing can substantially affect both fragility and vulnerability
curves, being therefore recommended that capacity curves can be derived based on
directly acquired data on local building stock, using available nonlinear commercial
analysis programmes (D’Ayala and Meslem 2013).
Moreover, one of the current challenges of this research topic is the harmonisation of
fragility curves, for the purpose of comparing different existing methodologies and
resulting curves. The harmonisation of fragility curves was addressed in the previously
mentioned SYNER-G project (Pitilakis et al. 2014), by means of a tool developed for the
storage and management of fragility curves. In order to accomplish this harmonisation,
the same intensity measure types, the same number of limit states and the same
building typology are needed.
Bearing in mind the exposed and for the sake of simplicity, the literature review herein
presented include the most relevant methodologies and respective fragility curves
developed within the European context, with particular focus to those developed for the
Mediterranean bordering countries. As already mentioned, these methodologies can be
grouped in four main approaches: analytical; empirical; expert judgment elicitation, and
hybrid approaches. In the following sections a very brief overview of each one of these
approaches is given, as the focus of this report are the analytical approaches, which will
be further detailed in the next Chapter 3.

2.1. Analytical approaches


Analytical approaches are commonly used for the construction of fragility curves due to
the way they address the problem of seismic vulnerability in structural engineering
terms, defining a direct relationship between construction characteristics, structural
response to the seismic action and damage effects (Rossetto et al. 2013). Within this
approach, most of the methodologies available in literature rest on two main and distinct
procedures: the correlation between acceleration or displacement capacity curves and
spectral response curves, as the well-known HAZUS or N2 methods (FEMA 2003a; Fajfar
2000), and the correlation between capacity curves and acceleration time histories, as
proposed in Rossetto and Elnashai (2005).
One of the first steps when conducting studies based on analytical approaches is the
definition of the sample or population of buildings that a given fragility curve will
represent. In this sense, the GEM Foundation has proposed a new building taxonomy
(Brzev et al. 2013) that includes information not only concerning the material and the

5
lateral load-resisting system itself, but also roof and floor systems, height, date of
construction, structural irregularity and occupancy class. This initiative is seen as a first
step towards the standardisation and harmonisation of building typologies that will
facilitate the comparison of fragility curves derived with different methodologies or to
different regions across the globe.
According to the scheme presented in Figure 1, the procedure for deriving analytical
fragility curves should follow the following path: defining of index buildings; defining of
components for response analysis and loss or repair cost estimation; selecting of the
model type; defining of the damage states at element and global level and defining the
analysis type and calculation of engineering demand parameters (EDPs) thresholds
between damage states, respectively from (A) to (E). Following this path one will be able
to run analysis and derive fragility curves (F), from which subsequently vulnerability
curves (G) can be generated by means of implementing appropriate damage-to-loss
functions.

Figure 1. Flow diagram for deriving analytical fragility curves.

Even though analytical methodologies involve the use of numerical or mechanical models
to evaluate the structural performance of buildings, allowing to take into account many
characteristics of the building stock in a direct, transparent, and detailed way, and also
to explicitly account for the uncertainties involved in the assessment procedure, they
naturally require a larger amount of detailed input data and a higher computational
effort, when compared with empirical methodologies (Del Gaudio et al. 2015). The most
relevant methodologies concerning the use of analytical methodologies for deriving
fragility curves are described and discussed in Chapter 3.

2.2. Empirical approaches


The seismic vulnerability assessment of buildings at large geographical scales has been
first carried out in the early 70’s, through the employment of empirical methodologies
initially developed and calibrated as a function of macroseismic intensities. This came as
a result of the fact that, at the time, hazard maps were, in their vast majority, defined in
terms of these discrete damage scales (Calvi et al. 2006). Therefore, empirical
approaches constituted for many years, the only reasonable and possible means of
developing seismic risk analyses at a large scale of assessment, based on the treatment
of post-earthquake damage observation data. The main drawbacks of this approach lies
precisely in the subjectivity on allocating each building to a damage state or in the lack
of accuracy in the determination of the ground motion affecting the region. Furthermore,
there are only a few dozen regions in the world where post-earthquake damage and
repair cost data has been collected from a number of buildings large enough to allow the
development of reliable fragility curves. The interdependency between macroseismic
intensity and damage and the limited or not homogeneous empirical data are commonly
identified as the main difficulties to overcome on the calibration process of empirical
approaches (Del Gaudio et al. 2015).
From the above, one can easily understand the need for treating uncertainty also when
dealing with this type of approaches. On one hand, aleatory uncertainty is introduced by
the characteristics of earthquakes and the resulting ground shaking and by the variation

6
of the seismic response of the buildings of a given typology. On the other hand,
epistemic uncertainty is introduced not only by small databases of often poor quality,
typically associated with errors introduced by missing data, biased sampling techniques,
misclassification errors or even data aggregated over large areas, but also by the
inability to account for the complete characteristics of the ground shaking in the
selection of measures of the ground motion intensity (Rossetto et al. 2013) as well as
the measurement error in the intensity measure levels at the required locations
(Rossetto et al. 2014).
The results of empirical methodologies are usually given in two different types: damage-
motion relationships such as damage probability matrices (DPM), and fragility curves.
While in the first type the conditional probability of obtaining a damage level due to a
ground motion of a given intensity is expressed in a discrete form, in the second
category, the probability of exceeding a given damage state is expressed as a
continuous function of the earthquake intensity.

2.3. Expert judgment elicitation approaches


Resorting to expert judgment becomes often inevitable in regions where there are no
empirical data or where the assets are difficult to model. Several authors contributed to
the establishment of a systematic approach to eliciting expert opinion (Winkler et al.
1992; von Winterfeld 1989). Cooke and Goossens (2000), for instance, provided formal
protocols, comprehensive procedures and guidelines on the elicitation process and
handling of such data in uncertainty analysis (Ouchi 2004). A common feature inherent
to expert judgment elicitation-based methodologies is the selection of a panel of experts,
which are recruited and trained for each particular study (Jaiswal et al. 2012). These
methodologies can differ significantly one from another on the collection of the expert
elicitation data process and its combination, but in general terms, those adopted in the
fragility literature, can be grouped into two main categories: mathematical and
behavioural methodologies (Clemen and Winkler 1999). In general, mathematical
methodologies to expert judgment elicitation are regarded to be more reliable,
reproducible and fair in aggregating expert opinions than behavioural methodologies
(Cooke 1991; Clemen and Winkler 1999; Jaiswal et al. 2012; Rossetto et al. 2014).
In mathematical methodologies, experts provide their estimate of an uncertain quantity
as a subjective probability without interacting among each other. Thus, after the
elicitation the estimates of the unknown quantity provided by each expert are combined
mathematically, typically using either a technique for weighting each expert’s estimates,
as in the case of Cooke (1991), or through the use of Bayesian statistics, as in the case
of Morris (1997). Such methodologies have been adopted in various financial and
environmental risk assessment exercises, for example in the estimation of volcanic risk
(Aspinall and Cooke 1998) or in seismic hazard estimation (Klügel 2011). Within the
fragility literature, Cooke’s classical method (Cooke 1991) was used by Jaiswal et al.
(2012) for the construction of fragility curves for the collapse limit state of reinforced
concrete (RC) and unreinforced masonry (URM) buildings, expressed as a function of
peak ground acceleration (PGA). In Cooke’s method, implemented in the free software
EXCALIBUR (Cooke 2001), each expert’s estimates of the unknown quantities of interest
are weighted according to their performance in answering a set of seed questions
considered relevant to the topic of interest, which are unambiguously worded and have a
unique and known numerical answer value. The responses from all experts are adopted
and the weightings of each expert are not revealed to participants in the process.
Behavioural methodologies, in turn, aim at producing some type of group consensus
among experts, who are typically encouraged to interact with one another and share
their assessments. One of the most well known behavioural methodologies is the Delphi
technique, which started to be developed in the early 1950s by Dalkey (1969). In this
method, experts are asked to anonymously judge the assessments made by other

7
experts within a panel. Each of the experts is then given a chance to reassess their
initial judgment based on the others’ review (Ouchi 2004). The process is the repeated
until a degree of agreement or consensus is reached between the expert estimates.
Later on, this method incorporated a self-rating mechanism, allowing experts to rate
their own expertise (Cooke 1991; Parenté and Anderson-Parenté 1987; Sackman 1975).
One of the best examples where a behavioural method has been used in building fragility
analysis is the ATC-13 (1985), in which a modified-Delphi method is applied to construct
Damage Probability Matrices expressing the likelihood of particular damage factors
(ratios of repair and replacement costs) being achieved over a range of Modified-Mercalli
Intensity (MMI) values, for 78 different building types in California (Jaiswal et al. 2012).
The main difference is that experts were asked to rate their experience and level of
confidence in their estimates, which were then used to weight their answers in the
combination of estimates of the value of interest by different experts (Jaiswal et al.
2012). The Nominal Group method is another well-known behavioural method, in which
experts are allowed to discuss their estimates directly with one another in a controlled
environment (Delbecq et al. 1975). According to Gustafson et al. (1973), this method is
considered more favourably than other group methodologies, particularly the Delphi
method.
Even though a group consensus method may help to identify experts’ errors and
misunderstandings during the elicitation process, there are no formal rules to apply in
order to reconcile differences when the consensus is difficult to achieve among different
experts (Ouchi 2004). Despite the literature on group consensus (Zahedi 1986;
Goicoechea et al. 1982; Eliashberg and Winkler 1981), the conformity induced by the
group interaction is still addressed as a major concern and motive of debate amid
researchers on the reliability of behavioural methodologies.
Mosleh et al. (1988) stressed out that the group interactive process could suffer, for
example, from the tendency for less confident experts to limit their participation, the
influence of dominant personalities and the tendency to reach speedy conclusions.
Likewise, Genest and Zidek (1986) pointed out that if unrestricted dialogue is allowed,
there might be room for strategic manipulation, bluffing, intimidating tactics and threats.
Cooke (1991), for example, has observed that more extreme probability estimates seem
to result from using the Delphi method, while Scheibe et al. (1975) pointed out that the
fact experts can see the responses of others may have a significant influence in swaying
expert’s initial judgments towards conformity to the majority of answers. Moreover,
authors such as Plous (1993) and Sniezek (1992), discussed issues related to group
polarization within this type of approach. It is generally agreed that mathematical
approaches yield more accurate results than do behavioural approaches in aggregating
expert opinions (Clemen and Winkler 1999; Mosleh et al. 1988).

2.4. Hybrid approaches


In addition to the main three above-mentioned approaches used to define fragility
curves for existing buildings, there is a forth type of approach, commonly designated as
hybrid, that combines different sources of data either analytical, empirical, or expert
judgment-based data. Hybrid damage probability matrices and fragility curves combine
usually post-earthquake damage statistics (empirical data) with simulated, analytical
damage statistics from a mathematical model of the building typology under
consideration (analytical data) (Calvi et al. 2006). Although they are actually rather
sparse, hybrid approaches can be particularly interesting either when there is a lack of
damage data at certain intensity levels for the geographical area under consideration or
when the calibration of analytical models needs to be carried out. Moreover, the use of
observational data reduces the computational effort that would be required to produce a
complete set of analytical fragility curves or DPMs (Calvi et al. 2006). One of the first
hybrid methodologies was proposed by Kappos et al. (1998), in which the damage data

8
used for the generation of the fragility curves was derived from a combination of
analytical simulations and post-earthquake surveys. In this work, nonlinear dynamic and
static analyses for all of the existing RC building typologies in Greece have been carried
out in order to extrapolate statistical data to PGA and/or spectral displacement values for
the cases in which no datasets were available, evidencing the potential of such type of
approaches (Kappos et al. 1998).
A hybrid method combining analytical, empirical and expert judgment-based data, was
used by the EERI-WHE study group in support of the PAGER project (Jaiswal and Wald
2010). This method aimed at defining the proportion of collapses for a given building
type given a shaking intensity expressed in the EMS-98 Intensity scale (Grünthal 1998)
for several nations worldwide where voluntary experts could be contacted. This exercise
highlighted the difficulty not only in correlating the expected percentage of damaged
buildings to a specific intensity scale, with which not all experts were familiar with, but
also in the definition of collapse. Given the large standard deviation of some of the
results obtained, Bayesian updating by means of empirical single event reported data
was used to improve experts’ forecast, as reported in D’Ayala et al. (2010) and Jaiswal
et al. (2011). Several improvements to this method were proposed by these authors,
such as the incorporation of uncertainty in the shaking intensity at which collapse
probability estimates are assigned, the improvement of the elicitation process, the
quantification of uncertainties associated with experts, the careful consideration of
variability in construction practices and building codes adopted worldwide when
comparing expert judgments for the same building types (Rossetto et al. 2014).

9
10
3. Review of analytical fragility assessment methodologies

This Chapter deals with the most relevant features concerning the construction of
fragility curves exclusively derived from analytical methodologies. Moreover, a
comprehensive review of some of the most well-known analytical fragility assessment
methodologies suitable to the European building stock is carried out aiming to contribute
to a better understanding of such methodologies, highlighting and commenting on their
main aspects, as for indicating the advantages and drawbacks of different methodologies
to support further decision making processes. Naturally, the following review has not
included all the existing procedures in literature since some of them are not applicable or
suitable to the European context, as for example (Erberik and Elnashai 2004; Jiang et al.
2012; Liel et al. 2009). However, the most important and widely used methodologies are
herein addressed.
Recent developments and technological improvements noted in the field of seismic
hazard assessment, such as the derivation of seismic hazard maps in terms of spectral
ordinates as opposed to macroseismic intensity or PGA, are seen as one of the main
reasons for which analytical methodologies have only been recently proposed for use in
risk and loss assessment (Calvi et al. 2006). Moreover, the use of such detailed and
transparent loss models allows detailed sensitivity studies to be undertaken, to which a
valuable insight on how much the loss results depend on the models, data, uncertainties
and assumptions employed can be given. Nevertheless, there is a general agreement
among the research community on the need to ensure an adequate balance between the
benefits of using analytical methodologies and the increased amount resources required
to construct an earthquake loss model based on such analyses. The capability of
numerical models to accurately predict the response of real structures, the accuracy in
transforming numerical indices of damage into actual damage of real structures, the
capability of accounting for human errors in the design and construction of buildings,
which are often the main causes of catastrophic collapses and the need to extend the
results obtained for few reference models to a large class of structures are some of the
issues stressed out by Calvi et al. (2006) that still need to be overcome.
Figure 2 illustrates the distribution of all the existing analytical methodologies considered
in this report, from where it is clear that these methodologies are distributed across the
Mediterranean Sea, the most seismic prone region in Europe. The number of
methodologies carried out in Italy and Turkey clearly stand out with respect to the
remaining countries, representing slightly more than 60% of the total number of
methodologies reviewed in this report. This fact can be explained by the significant
losses suffered by these two countries in the last decades from seismic events such as
the April 6, 2009 L’Aquila and the May 20, 2012 Emilia-Romagna earthquake in Italy, or
the August 17, 1999 Izmit and the October 23, 2011 Van earthquakes in Turkey. These
events with magnitudes, Mw, ranging from 5.8 up to 7.6, caused extended damage to
the building stock, triggering the opportunity for using empirical data for validating and
calibrating analytical methodologies.
According to D’Ayala and Meslem (2013), analytical fragility assessment methodologies
are commonly based on two main components, namely the structural response-to-
damage state functions, which are the product of two independent procedures, structural
and damage analysis, and to which a certain level and type of uncertainty should be
expected and accounted for by users when performing seismic risk assessments, and the
ground motion intensity-to-structural response functions. Consequently, the level of
uncertainty depends upon the simplifications and assumptions considered to lower
resource consumption to convenient levels, particularly in what regards data collection
and computational efforts. It is common in risk analysis and engineering safety problems
to distinguish between uncertainty that reflects the inherent natural variability of the
outcome of a repeatable experiment (aleatory uncertainty) and modelling and statistical
uncertainty due to limited or imperfect knowledge (epistemic uncertainty).

11
Figure 2. Distribution of the examined analytical fragility assessment methodologies by country:
Italy (12); Turkey (11); Greece (8); Portugal (3); Romania (2); Switzerland (2); Former Yugoslav
Republic of Macedonia (1); Slovenia (1) and Spain (1).

It is likely that epistemic uncertainties will be reduced as the understanding of the


variables increases, as for example, through the collection and analysis of additional
data and the development of improved predictive models (Faber and Stewart 2003).
Modelling uncertainty arises from the fact that every model, conceptual or mathematical,
is unavoidably a simplified representation of the reality and thus derives from the lack of
knowledge on the geometrical and mechanical characteristics of the structure and/or by
the limitations of the model to correlate these characteristics and to perform structural
analyses. Statistical uncertainty is related to the statistical evaluation of results of tests
or observations. It may result from a limited number of observations or data, neglecting
systematic variations of the observed variables or possible correlations between these
variables.
Generally, most of the available methodologies to develop analytical fragility curves
include the following steps concerning uncertainty assessment: identification of sources
of uncertainty on capacity, demand and damage thresholds definition; quantification of
those uncertainties and their modelling when constructing fragility curves. With respect
to the structural analysis phase, uncertainties related to capacity and demand modelling
are typically accounted for in the estimation of the building performance, as shown in
Figure 3.
Uncertainty in capacity is directly affected by the parameters related not only to the
building’s structural characteristics, which include the mechanical properties, structural
details, geometric configuration and dimensions, but also to the mathematical model
used to compute the building’s structural capacity, which in turn depends on the
detailing level and completeness of the numerical model, and on the performance criteria
considered. Uncertainty in demand is introduced by the quality of the ground motion
(seismic demand) representation through real ground motion records or code-based
spectra that need to capture the record-to-record variability, i.e. the variability of the
seismic source mechanism, attenuation and site effects on the seismic event (FEMA
2003b; NIBS-FEMA 2003; ATC 2011).
The damage thresholds modelling, included in the damage analysis phase, is affected by
the definition of the different global limit states, the choice of the damage model and its
consistency with the type of analysis, the damage indicator used to represent the
damage states of a structure and the correlation with the chosen intensity measure

12
(FEMA 2003a; NIBS-FEMA 2003). It is perhaps one of the most challenging sources of
uncertainty to be overcome, as there is a lack of clear guidance in literature concerning
this topic, beyond some qualitative description of observed damage (D’Ayala and Meslem
2013).
For the time being, the existing expressions for the calculation of demands and
capacities are mostly defined at the element rather than at the global level. Several
definitions have been implemented in guidelines and codes for the estimation of the
global damage states through the observation of the progression of local damage at
elements. According to Eurocode 8 (CEN 2004), two limit states are defined in relation to
the fundamental performance requirements and compliance criteria for structures within
seismic regions:
 Ultimate limit states (ULS), associated with collapse or with other forms of
structural failure which might endanger the safety of people;
 Damage limitation states (DLS), associated with damage beyond which specified
service requirements are no longer met.

The fragility analysis phase includes the choice of the fitting and sampling methods, the
selection of models to express the fragility curves and the construction process itself,
taking into account the uncertainties considered and measured in the previous structural
and damage analysis phases (FEMA 2003a; Wen et al. 2004; Pagnini et al. 2011; ATC
2011).

Figure 3. Main components and phases considered in analytical fragility assessment methodologies
and associated uncertainties.

13
3.1. Structural system
Structural walls, which can be taken as vertical planar elements that resist not only to
gravity loads but also to horizontal forces, provide stability to a building during an
earthquake. They may be monolithic as in the case of reinforced concrete walls, or be
diversified as in the case of masonry walls. The majority of the studies found in literature
that considered this structural system are related to the assessment of masonry structures,
as the example illustrated in Figure 4 (left) of a masonry structural wall system and (right)
the corresponding three-dimensional assemblage of all the structural walls, modelled
through a macro-element approach using the 3Muri ® software (STADATA 2011), widely
used for performing nonlinear static (pushover) analysis of masonry structures. Relevant
studies were carried out by Barbat et al. (2006), Erberik (2008), Borzi et al. (2008c), who
adapted the SP-BELA (Simplified Pushover-Based Earthquake Loss Assessment) method to
masonry buildings, once it was originally developed for the determination of the seismic
vulnerability of RC buildings, and Ceran (2010), for example.

Figure 4. Example of a masonry structural wall system (left) and the corresponding three-dimensional
assemblage of all the structural walls (right), modelled through a macro-element approach using the
3Muri® software (Ferreira et al. 2015).

With respect to reinforced concrete structures, moment-resisting frames, infilled frames and
dual frame-wall systems are the most recurrent structural systems considered in fragility
assessment studies. Moment-resisting frame systems, as shown in Figure 5 (left), are
widely used to resist lateral loads through axial forces, shears, and moments in their beams,
columns and rigid beam-to-column connections. Their strength and ductility arise from the
combination of concrete and reinforcing steel that resist compressive and tensile forces
respectively. However, as pointed out by Barbat et al. (2006), in some particular cases as in
the city of Barcelona, Spain, most of the reinforced concrete buildings are not moment-
resisting frames but typically column and slab buildings in their waffled slab floor version,
which typology is highly not recommended for seismic prone areas due to their low ductility.
The majority of the methodologies herein analysed were developed for moment-resisting
frame buildings not only because they cover a large fraction of the European reinforced
concrete building stock, but also because the numerical representation of this particular
typology results in general more reliable due to its inherent simplicity when compared to
both infilled and dual frame-wall systems.
Reinforced concrete infilled frame systems can be defined as a framework of beams and
columns in which some bays of frames are infilled with masonry walls that may or may not
be mechanically connected to the frame. Due to the great in-plane stiffness and strength,
infill walls do not allow beams and columns to bend under horizontal loading, changing the
structural performance of the frame. Thus, due to horizontal earthquake-induced actions,
diagonal compression struts are formed in infills so that the structure behaves as a braced

14
frame rather than a moment frame. Moreover, infill walls can fill partially or completely the
height and or length of a bay. Soft-storey collapse mechanisms are introduced when infill
walls are discontinued at a determined storey. It has been experimentally documented that
the presence of infill walls significantly increases the strength, stiffness and seismic energy
dissipation capacity of buildings (Kappos et al. 2003; Tsionis et al. 2014). Relevant
methodologies concerning this particular issue were included in this review, namely those
developed by Kappos et al. (2003), Akkar et al. (2005), Kappos et al. (2006) and Del
Gaudio et al. (2015). Further comments concerning its influence and its consideration in
structural modelling will be addressed in the following section.
In reinforced concrete dual frame-wall systems, the lateral load-resisting structure
comprises both moment-frames and shear walls that act together in the same direction.
These walls are usually not perforated by openings and can be found around the staircases,
as demonstrated in Figure 5 (right), elevator shafts and at the perimeter of the building.
When wisely designed, the presence of such walls may have a positive effect not only on the
seismic performance of the frame structure but also on preventing soft-storey collapse
mechanisms. However, the slenderness of walls in dual frame-wall systems should be
carefully considered. Depending on the share of shear resistance at the base of the walls
and frames, Eurocode 8 (CEN 2004) defines three systems – frame-equivalent dual, wall-
equivalent dual and wall – and particular design rules for each of them. It is worth referring
that the studies by Dumova-Jovanoska (2000), Kappos et al. (2006), Kappos and
Panagopoulos (2009) and Bilgin (2013) were carried out to understand the differences in
terms of seismic fragility between moment-resisting frames, infilled frames and dual frame-
wall systems.

Figure 5. Examples of reinforced concrete moment-resisting frame (left) and dual frame-wall
structural systems (right) (Dumova-Jovanoska 2000).

3.2. Components for response analysis


As becomes clear from the previous section, two groups of components can be considered in
analytical fragility and vulnerability assessment methodologies: structural and non-
structural components. Structural components are the main elements that directly

15
contribute to the seismic response of the structure and therefore are considered in both
mathematical and loss estimation models. Moreover, the linear and nonlinear behaviour of
structural components must be defined and explicitly simulated in the analysis to be
undertaken. Even though horizontal components such as roofs and floors systems have a
structural role, they are hardly explicitly modelled for the sake of modelling simplification,
being their effect usually simulated by introducing specific constraint conditions or
assumptions. Non-structural components may contribute to the response behaviour of the
structure, as in the case of masonry infill walls in RC buildings, or be considered only in loss
estimation modelling, in case they represent a significant weight over the building
construction cost (D’Ayala et al. 2014; Dolšek and Fajfar 2008).
Masonry infill walls are made of masonry bricks or blocks, varying in specific weight,
strength and brittleness, depending on age and quality of construction, and are numerically
represented by means of an equivalent diagonal strut model, as demonstrated in the
following Figure 6.

Figure 6. Numerical representation of infill masonry walls through equivalent diagonal strut model
(Amato et al. 2008).

Non-structural components are only taken into account for the case of fragility assessment
of RC buildings, as clearly demonstrated in the above-mentioned studies of Kappos et al.
(2003), Akkar et al. (2005), Kappos et al. (2006) and Del Gaudio et al. (2015). Although
the consideration of the contribution of masonry infill walls in the analysis model has been
highly recommended due to its significant influence on the lateral resistance of a reinforced
concrete frame, as experimentally pointed out for example in Fiorato et al. (1970), Klingner
and Bertero (1976), Bertero and Broken (1983) and Mehrabi et al. (1994), still few studies
actually account for their effect on the seismic response of RC buildings.
A major reason for this tendency is related to the difficulty on determining the type of
interaction between the infill and the frame. D’Ayala and Meslem (2013) have pointed out
that the exclusion of infills’ contribution may lead to significant bias in resulting fragility
curves. Moreover, as stressed out in Figure 7, when masonry infilled RC buildings are
modelled as moment-resisting (or bare) frame structures, fragility curves tend to achieve
greater lateral displacement capacity for all damage levels, whereas, for the case when the
infilled frame model is considered, the building is found to be more vulnerable.

16
Figure 7. Comparison between fragility curves for 4-storey RC buildings, with and without considering
the modelling of masonry infill walls (D’Ayala and Meslem 2012).

3.3. The selection of the analysis type


There are different options available for the choice of mathematical modelling and type of
analysis for structural assessment. As mentioned previously, uncertainties related to these
models are directly reliant on the level of complexity desired to conduct the fragility
assessment. In view of the above, analysis can be grouped, in decreasing order of
complexity, as follows:
 Nonlinear Dynamic (NLD) analysis, which requires a set of ground motion records to
perform dynamic response history analysis of a mathematical model;
 Nonlinear Static (NLS) analysis based on the use of a (first-mode or other) load
pattern to perform a pushover analysis of a structure, and then fit the resulting
capacity curve with an appropriate back-bone curve response model, e.g. elastic-
plastic, elastic-plastic with residual strength, quadrilinear, etc;
 Nonlinear Static analysis based on Simplified Mechanical Models (SMM-NLS), in which
the capacity curve is obtained through simplified analytical methodologies that do
not require finite element modelling;
 Linear Static (LS) analysis, which can be a modal response spectrum analysis or one
using the lateral force method.

The level of detailed modelling of each structural component will depend on the choice of
analysis type selected. For all types of analysis, D’Ayala et al. (2014) suggest that the user
should use median values of structural characteristics-related parameters when defining the
component behaviour. When median values are not available, mean values should be used.
The user should make sure to simulate all possible modes of component damage and failure
(axial, flexural, flexure-axial interaction, shear, and flexure-shear interaction), or P-Delta
effects, for example. Relevance guidance may be found in design standards, such as the
Eurocode 8 (CEN 2004).

3.3.1. Nonlinear Dynamic Analysis


From the methodologies considered in this report it was possible to identify different
approaches to perform nonlinear dynamic analysis for the structural assessment of
buildings: conventional approaches; incremental dynamic analyses (IDA), and hybrid

17
approaches. Some authors, as for example Akansel et al. (2012), Hancilar and Çaktr (2015)
or Özer and Erberik (2008), carried out a conventional approach by performing nonlinear
time-history (dynamic) analyses of RC structures, using software such as ANSYS ® (ANSYS
2005), IDARC-2D® (IDARC-2D 1996) or OpenSEES® (2015). In these studies, the seismic
demand was obtained for different sets of ground motion, the capacity was determined in
terms of limit states and the corresponding fragility curves were obtained by assessing the
probability of reaching or exceeding each limit state for different levels of ground shaking
intensity. However, in Dumova-Jovanoska (2000), damage curves were derived by means
of a damage index, which was then plotted versus the Modified Mercalli Intensity.
Incremental Dynamic Analysis (IDA) is often perceived as the dynamic equivalent to a
pushover analysis, and has been recommended in ATC-63 (FEMA 2008) and ATC-58 (FEMA
2012) technical reports. Even though the IDA approach can be implemented to any building
typology to estimate different median capacities, it entails the definition of a complete
hysteretic behaviour of the materials and the repetition of the analysis for a large number of
acceleration time histories, to which a significant time consuming computation process
might be associated, depending on the level of complexity and material type considered.
Moreover, it is fundamental to ensure that the model is consistent with the type of analysis
to be carried out and that sufficient model complexity is retained. To this end, it is
necessary to define hysteretic curves for structural and possibly also for non-structural
elements, to use median values for structural characteristics-related parameters, to
simulate all possible modes of component damage and failure, and to define permanent
gravity actions (D’Ayala et al. 2014). Regarding the minimum number of ground motions
that should be used to provide stable estimates of the median capacity, it has been reported
that the use of 11 pairs of motions (i.e. 22 motion set, including two orthogonal
components of motion), should be sufficient for the purpose of this procedure (D’Ayala et al.
2014). According to this method, nonlinear response history analysis is performed for
increasing intensity of the ground motion, until either global dynamic or numerical instability
occurs in the analysis. The main output of IDA analyses is a curve that plots maximum
storey drift at any level of building or other EPD, versus spectral acceleration, Sa(T), or
other intensity measure, as demonstrated in the example of Figure 8. The collapse capacity
point of IDA curves is defined by the point from which the curve starts to become constant
over the intensity measure values.

Figure 8. IDA curve with the identification of yielding and collapse points (Kirçil and Polat 2006).

From the reviewed literature there are only a few authors that have conducted IDA, namely
Kirçil and Polat (2006) and Pitilakis et al. (2014b). In the first, IDA analyses were performed
to determine the yielding and collapse capacity of the sample buildings in terms of elastic
pseudo-spectral acceleration, under the effect of twelve artificial ground motions. Fragility
curves were then derived to express the probability of structural damage due to

18
earthquakes as a function of ground motion intensity, expressed in terms of 5% damped
pseudo elastic spectral acceleration (Sa) and assuming that the fragility curves can be
expressed in the form of two-parameter lognormal distribution functions (Kirçil and Polat
2006). Based on this assumption, the cumulative probability of occurrence of the damage,
equal or higher than damage level Di, was defined as:

(Eq.1)

where Φ is the standard normal distribution, X is the lognormal distributed pseudo elastic
spectral acceleration, λ and ζ are the mean and standard deviation of ln X. Here lognormal
probability paper was used for obtaining mean and standard deviation of ground motion
indices for each damage level. Lognormal probability paper was used for obtaining mean
and standard deviation of ground motion indices for each damage level. This method is
based on plotting ln X versus the corresponding standard normal variable on a lognormal
scale and performing a linear regression analysis to determine the mean and standard
deviation of ln X for each damage level Di (Gündüz 1996).
In Pitilakis et al. (2014b) in turn, two-dimensional incremental dynamic analyses were
performed to assess the seismic performance of the uncorroded and corroded RC moment-
resisting frame structures, designed which different seismic codes. The time-dependent
fragility curves were then derived in terms of the spectral acceleration at the fundamental
mode of the structure and the outcropping peak ground acceleration for the immediate
occupancy and collapse prevention limit states.
Finally, hybrid approaches within NLD analysis are often used, in which different types of
analyses are combined. Rota et al. (2010) used nonlinear static (pushover) analyses to
define the probability distributions of each damage state and nonlinear dynamic analyses to
determine the probability density function of the displacement demand corresponding to
different levels of ground motion. Casotto et al. (2015) took advantage of pushover analysis
to establish a number of damage limit states and then performed nonlinear dynamic
analysis to compare the maximum demand with the limit state capacity to allocate the
structure in a damage state.
Silva et al. (2013) stressed out that the technique by which the actual dynamic
phenomenon is reproduced when employing nonlinear dynamic analysis, i.e., applying an
acceleration time history at the base of the structure, leads in theory to more accurate
results, being hence one of the main advantages of this type of approaches for deriving
fragility curves. However, the intrinsic modelling complexity (e.g. hysteretic response
models, equivalent viscous damping) combined with the heavy computational effort, is often
impractical, thus favouring the employment of simpler methodologies, comprising nonlinear
static analysis.

3.3.2. Nonlinear Static Analysis


Nonlinear Static Analysis (NLS), or simply pushover analysis, has become a very popular
analysis tool for the seismic performance evaluation of both existing and new structures
under permanent vertical loads and gradually increasing lateral loads, with the latter
representing approximately the earthquake-induced forces. The capacity curve is the output
from NLS analysis and it represents a plot of the total base shear versus top displacement,
which is then fitted to a bilinear (elastic-plastic), as illustrated in Figure 9 (left), or multi-
linear (e.g. elastic-plastic with residual strength) curve. Moreover, as this analysis is carried
out up to failure, it enables the determination of collapse load and ductility capacity. The

19
derivation of a capacity curve by use of nonlinear static (pushover) analysis, does not
directly account for the specific seismic motion causing the collapse, as the dynamic
characteristics are not taken into account in the analysis. Moreover, the consideration of
higher mode effects once a local mechanism is formed and issues related to cumulative
damage have been pointed out as some of the main drawbacks of NLS analysis. In this
sense, a new extension to the previously mentioned conventional pushover analysis has
been proposed in recent years, the so-called Static Adaptive Pushover Analysis (Antoniou
and Pinho 2004; Papanikolaou et al. 2005). The main difference is that when performing a
SAPA, the lateral load distribution is not kept constant but it is rather continuously updated
during the analysis, according to the modal shapes and participation factors derived by
eigenvalue analysis carried out at each analysis step. Figure 9 (right) illustrates a set of
capacity curves derived for the randomly generated RC frames by adopting a displacement-
based adaptive pushover (Silva et al. 2013).

Figure 9. Example of (bilinear) capacity curves derived from conventional pushover analysis (left) and
those obtained from using a displacement-based adaptive pushover (right), developed by Ferreira et
al. (2015) and Silva et al. (2013), respectively.

Due to its ability to update the lateral load patterns according to the constantly changing
modal properties of the system, SAPA overcomes the intrinsic weaknesses of fixed-pattern
pushover analysis and provides a more accurate performance-oriented tool for structural
assessment, providing better response estimates than existing conventional approaches,
especially in the presence of significant strength or stiffness irregularities (Papanikolaou et
al. 2005). It is worth to recall that SAPA approach has been used in nonlinear studies
carried out both in masonry buildings, as in Lourenço et al. (2011) and in RC buildings, as in
Chaulagain et al. (2013). Furthermore, Abbasnia et al. (2013) has pointed out that SAPA
analyses can capture the results of IDA analysis with a reasonable accuracy. However, the
excessive force concentration at the locations of the structure where the damage first
develops, the combination of the modal contributions, the updating procedure of the lateral
load vector and the computation time were identified as major drawbacks of this approach
(D’Ayala et al. 2014).

3.3.3. Simplified Mechanism Models


A number of simplified analysis methodologies to derive fragility curves exist in the
literature. These methodologies are based on assumptions that allow the derivation of
vulnerability or fragility curves without running a full non-linear analysis. Although these
methodologies might have the advantage of allowing analysis of large populations of

20
buildings with relatively little calculation burden, the reliability of the results is highly
dependent on the expertise of the operator in choosing the underlying simplifications and
assumptions. Such methodologies are more common in the assessment of masonry
structures and can be subdivided in methodologies that yield fragility curves based on
lateral capacity and those based on displacement.
Collapse mechanism-based methodologies such as the VULNUS (Bernardini et al. 1990), or
FaMIVE (D’Ayala and Speranza, 2003) are procedures based on collapse-load factors or
collapse multipliers associated to different modes of collapse, intentionally developed for
assessing the fragility of masonry structures. The method proposed by Cosenza et al.
(2005) uses a similar approach for the assessment of reinforced concrete buildings. In
VULNUS and FaMIVE the aleatory variation associated with geometry and materials is
accounted for by applying the procedure to a large number of buildings at a particular
geographical location. Epistemic uncertainty related to the various parameters that enter
the calculation is accounted for by applying to each parameter a confidence range variable
between 0.10 and 0.30, while in VULNUS fuzzy membership classes are defined.
Simplified displacement-based methodologies use an idealised single degree of freedom
nonlinear oscillator to derive a bilinear capacity curve that accounts for the post-peak
behaviour of the structure. An application for RC structures was proposed by Pinho et al.
(2002) and improved by Crowley et al. (2004), giving rise to the DBELA method. A similar
application was developed for URM, the so-called SPBELA method (Borzi et al. 2007; Borzi
et al. 2008c). Both methodologies define the damage state limit thresholds of structures at
the level of the equivalent single element. Subsequently, an idealised bilinear capacity curve
to intersect then with a response spectrum is defined in order to define representative
performance points and ultimately derive fragility curves. A different approach is used by
D’Ayala (2005) to extend the applicability of the mechanism approach in FaMIVE within a
displacement-based assessment framework. A different capacity curve is associated to each
failure mechanism, parameterised in terms of total drift and ductility to define fragility
curves. The quality and reliability of these procedures is dependent on the validation of the
idealised capacity curves and definition of the associated limit states by experimental or
empirical evidence.

3.3.3.1. FaMIVE method


The FaMIVE (Failure Mechanism Identification and Vulnerability Evaluation) method,
estimates the building performance both in terms of base shear and deformation capacity
and identifies the most suitable strengthening and repair intervention by considering the
possible collapse mechanisms which can occur given geometry, materials, loading conditions
and constraints such as ring beams or tie-rods (D’Ayala and Speranza 2003). Both in-plane
and out-of-plane failure mechanisms are considered, being the latter illustrated in Figure
10. Using the concepts of limit state analysis, this method correlates collapse mechanisms
to specific construction features of the external bearing walls forming a masonry building
(D’Ayala et al. 1997). The analysis quantifies the collapse load factor (as a percentage of
gravity acceleration) associated with each mechanism so as to determine a lower bound of
the level of shaking which will trigger the onset of a specific failure mechanism (D’Ayala et
al. 2005). This method uses a nonlinear pseudo-static structural analysis by means of a
variant of the N2 method (Fajfar and Gašperšič 1996), included in Eurocode part 3 (CEN
2004), in which a degrading pushover curve is used to estimate the performance points, as
it has been described by Dolšek and Fajfar (2004).
The fragility curves for different limit states are then obtained by computing the median and
standard deviation values of the performance point displacements for each index building of

21
a given sample and by deriving the corresponding lognormal distributions (D’Ayala et al.
2014).
Moreover, this method can be applied to medium-size samples of buildings without requiring
a detailed analysis of the geometric, typological and structural parameters that characterize
the analysed buildings. The data that need to be collected focus on the parameters that
directly influence the structural performance of buildings. In this way, the method minimizes
the surveying time and the need for pre-existing information while providing an analytically
based vulnerability assessment (D’Ayala et al. 2005).

Figure 10. Out-of-plane failure mechanisms of the FaMIVE method (D’Ayala and Speranza 2003).

The method has been applied to estimate the performance of buildings in several locations
worldwide such as the Alfama District of Lisbon (D’Ayala et al. 1997), the Marche region in
Italy (D’Ayala and Speranza 2002), the historical center of Vittorio Venetto (Bernardini et al.
2008) and San Giuliano di Puglia (Indirli et al. 2004), Nepal (D’Ayala, 2004), India (D’Ayala
and Kansal, 2004), the Fener–Balat sample in Istanbul (D’Ayala and Yeomans 2004),
L’Aquila following the 2009 earthquake (D’Ayala and Paganoni, 2011), and recently in the
historical centres of Bovec and Ljubliana in Slovenia (Bosiljkov et al. 2012) and in the
Casbah of Algiers (Novelli and D’Ayala, 2014). Thus, within the FaMIVE database built over
the past years, capacity and fragility curves are available for various unreinforced masonry
typologies, from adobe to concrete blocks, for a number of reference typologies and in a
number of regions of the world (D’Ayala and Kishali 2012, D’Ayala 2013).

3.3.3.2. DBELA method


The Displacement-Based method for Earthquake Loss Assessment (DBELA) method is a
simplified nonlinear static analysis method for the seismic risk assessment of buildings, built
upon the urban assessment methodology proposed by Calvi (1999), and employed to
estimate the nonlinear capacity of thousands of RC frames randomly generated and the
corresponding demand from a large set of ground motion records. The fact that several
synthetic buildings and ground motion records are used in the calculations allows the
consideration of the material and geometrical uncertainties, as well as (to some extent) the
record-to-record variability (Silva et al. 2013). The principles of structural mechanics and
seismic response of buildings are used to estimate the seismic vulnerability of different
classes of buildings (D’Ayala et al. 2014), namely RC and URM buildings (Glaister and Pinho

22
2003; Restrepo-Velez and Magenes 2004; Crowley et al. 2004; Borzi et al. 2008a; Borzi et
al. 2008b). The method is further developed to derive fragility curves for building classes
considering their global and local vulnerabilities and their corresponding mechanical models,
defined completely by secant vibration period, viscous damping and limit state displacement
capacities. Each limit state marks the threshold between the levels of damage that a
building might withstand, usually described by a reduction in strength or by the exceedance
of certain displacement or drift levels. Once these parameters are obtained, the
displacement capacity of the first limit state is compared with the respective demand. If the
demand exceeds the capacity, the next limit states need to be checked successively, until
the demand no longer exceeds the capacity and the building damage state can be defined.
If the demand also exceeds the capacity of the last limit state, the building is assumed to
have collapsed. According to Silva et al. (2013), this procedure, which flowchart is
presented in Figure 11, provides a good balance between computational efficiency and
reliability, allowing a quick and simple assessment of the physical vulnerability of many
different building typologies.

Figure 11. Flowchart of the DBELA method (Silva et al. 2013).

3.3.3.3. SP-BELA method

The SP-BELA (Simplified Pushover-Based Earthquake Loss Assessment) method combines


the definition of a pushover curve using a simplified mechanic-based procedure, similar to
that proposed by Cosenza et al. (2005), to define the base shear capacity of the building
stock, with a displacement-based framework similar to that in DBELA, such that the
vulnerability of building classes at different limit states can be obtained. The main
component of the methodology involves the definition of the capacity of a population of
buildings based on a prototype structure, which is carried out using simplified pushover
analysis to obtain the collapse multiplier Borzi et al. (2008b). In Figure 12, the flowchart
behind the SP-BELA method is illustrated.

23
Figure 12. Flowchart of the SP-BELA method (Borzi et al. 2008b).

According to Borzi et al. (2008b), which have conducted a comparison between DBELA and
SP-BELA methodologies, pointed out that given its more simplified nature, DBELA requires
an assumption on the likely failure mode of structures, beam-sway or column-sway, to be
made a priori. Instead, in SP-BELA method such assumption is not required, since the
failure mechanism is computed by the algorithm itself, as discussed in Borzi et al. (2008b).
Moreover, the DBELA methodology requires a combination of the beam-sway and column-
sway curves depending on the proportions of the building stock that are assumed to fall into
either of these two groups. However, such assumptions are not necessary with the SP-BELA
procedure and the proportion of each mechanism is implicitly accounted for in the
generation of the vulnerability curves. It is reassuring to note from observing Figure 13,
that the SP-BELA curve is closer to the DBELA column-sway curve than the beam-sway
curve as the buildings considered herein have not been designed considering capacity

24
design principles and thus the formation of plastic hinges in the columns, as opposed to the
beams, is expected Borzi et al. (2008b).

Figure 13. Vulnerability curves for 4-storey buildings using the SP-BELA and DBELA methodologies for
Limit State 2 (left) and Limit State 3 (right) conditions (Borzi et al. 2008b).

3.3.3.4. VULNUS method


The VULNUS method allows estimating the seismic vulnerability of a single building using
the fuzzy-set theory and the definition of collapse multipliers (Bernardini et al. 1989;
Bernardini et al. 1990). Developed by researchers of University of Padua in the second half
of the ‘80s, this method was recently modified according to the Italian Seismic Code and
written in Visual Basic programming language (Munari et al. 2009). The approach is based
on building survey, in order to collect geometrical and structural information, handled
through qualitative judgment (Florio 2010). VULNUS estimates the global vulnerability of
regular (both in plan and in height) masonry structures with a limited number of storeys. It
applies either to single buildings or building aggregates. The major limitation of this method
is the static treatment of the dynamic seismic action.

3.4. The selection of the model type


According to international scientific literature, three types of models are used for fragility
analysis, offering distinct choices of structural detailing. In decreasing order of complexity,
they are:
 Multi-Degree-of-Freedom (MDoF) model (two/three-dimensional elements);
 Reduced MDoF model (two-dimensional elements);
 Single-Degree-of-Freedom (SDoF) model.

Even though SDoF models are not the most accurate, the simplification to an equivalent
single-degree-of-freedom system is widely used when computing the performance point and
damage state of a structure in a number of methodologies for the determination of fragility
curves.
The employment of two-dimensional (planar) models for the derivation of fragility curves or
capacity curves has been quite widely employed in literature, for instance, in Erberik and
Elnashai (2004) for a 5-storey RC building, in UTCB (2006) for a 13-storey RC building in
Bucharest, or in Barbat et al. (2006) for a range of RC and URM in Barcelona with two to

25
eight storeys. The aim of using two-dimensional models is to reduce the computational
effort, especially when using nonlinear dynamic analysis. However it should be kept in mind
that the epistemic uncertainty associated with this modelling strategy can be significant,
especially for buildings with irregular geometries or with buildings with non-uniform
distribution of infills. Figure 14 shows the results from a study carried out by D’Ayala and
Meslem (2013) that compared the capacity curves of an RC frame structure with infills
obtained from a three-dimensional model with those obtained by the superposition of two-
dimensional models. It can be observed that there was a quite significant difference
between both procedures. By using two-dimensional models, the displacement
corresponding to first yielding of steel and first crushing of concrete member seems to be
overestimated. In addition, the peak loading capacity is underestimated by using two-
dimensional models. Moreover, despite the discrepancies in the estimated seismic response
of structures when employing different methodologies, as already shown by many authors
(Chopra and Goel 2000, Casarotti et al. 2009, Lin et al. 2004), it was concluded by Silva et
al. (2013) that these relevant differences are not necessarily reflected into the resulting
fragility curves, particularly for lower damage states. This occurs because often the focus is
not on the member level performance (in terms of bending moments or shear forces) of an
individual structure, but rather on the global damage state of the building.

Figure 14. Comparison of capacity curves obtained by the three-dimensional model with those
obtained by superposition of two-dimensional models (D’Ayala and Meslem 2013).

3.4.1. MDoF models


A multi-degree-of-freedom model of a structure is a detailed three-dimensional or two-
dimensional model that includes elements for each identified lateral-load resisting
component in the building, as columns, beams, infills, walls, etc. For the employment of
three-dimensional or two-dimensional element-by-element models, the identification of
building’s components, the definition of the non-structural elements, the determination of
the foundation flexibility, level of diaphragm action, and the definition of permanent gravity
actions (i.e. dead loads and live loads) is necessary. Preferably, the building should be
modelled as three-dimensional, as for instance in Akansel et al. (2012), Pitilakis et al.
(2014b) or Hancilar and Çaktr (2015). However, in some cases, the user may wish to use
two-dimensional (planar) models in order to reduce the calculation effort. This may be
acceptable for buildings with regular geometries where the response in each orthogonal
direction is independent and torsional response is not significant.

26
3.4.2. Reduced MDoF models
A reduced MDoF model, such as a stick model or a single-bay frame one, is a simplified two-
dimensional lumped stiffness-mass-damping representation of a building, in which each of
the N floors is represented by one node having three degrees of freedom, two translational
and one rotational to allow the representation of both flexural and shear types of behaviour
(D’Ayala et al. 2014). This representation is not suitable for irregular buildings or buildings
where torsional effects are anticipated. As appropriate component level modelling requires
advanced structural skills and it is a critical aspect of the vulnerability estimation, when data
is not very accurate or resources are modest, it might be worth to adopt a reduced MDoF
model employing storey-level, rather than component-level, characterization of mass,
stiffness and strength. It should be noted that substantial approximations are made in
adopting these models, in the determination of storey level mass stiffness and strength,
which assume either an average or homogenous behaviour of single components. This may
substantially affect the expected global failure modes. The advantage is that such models
can be analysed within a few seconds using either nonlinear dynamic or static methods. Of
course, such modelling techniques are only appropriate for structures having rigid
diaphragms without appreciable plan irregularities.
One example of reduced MDoF models is given by the study developed by Kappos and
Panagopoulos (2009), in which inelastic static and dynamic time-history analyses were
carried out for RC buildings using the SAP2000N (Computers and Structures Inc. 2002) and
the DRAIN-2000 (Kappos and Dymiotis 2000) codes, respectively. RC members were
modelled using lumped plasticity beam-column elements, while infill walls were modelled
using diagonal equivalent strut elements for the inelastic static analyses and shear panel
isoparametric elements for the inelastic dynamic analyses. Moreover, to make the analysis
process less time-consuming, two-dimensional models of these buildings were analysed, as
exemplified in Figure 15. The fact that two-dimensional models ignore effects such as
torsion due to in-plan irregularity is stressed out as its major drawback.

Figure 15. Configuration and dimensions of typical four-storey irregularly infilled RC frame building
(Kappos and Panagopoulos 2009).

3.4.3. SDoF models


A single-degree-of-freedom model is a simple one-dimensional (linear or nonlinear) element
representing the stiffness, mass, damping and ductility of the structure. This representation
is in general simplistic and assumes that higher vibration modes are not relevant to the
seismic response of the structure. Although a poor approximation of the real behaviour of

27
the structure is obtained with this type of model, SDoF equivalent models might be
employed when there are modest resources or the knowledge of the structural
characteristics is so poor that does not warrant the effort of detailed modelling. This
becomes possible by adopting mechanical models that are able to represent the dominant
response characteristics of specific structural typologies, in which the analytical formulation
is derived from the failure mechanisms and behaviour of structures observed during
experiments or real earthquakes. Thus, according to D’Ayala et al. (2014), each such model
should use analytic expressions or simple calculations to accurately simulate at least:
• The capacity curve of the structure;
• The first mode period and associated mass;
• The equivalent stiffness of the system.

There are several methodologies that use equivalent SDoF models for the numerical
representation of low- up to high-rise structures, as for example the above-mentioned
DBELA method. Moreover, the seismic performance-based assessment of typical masonry
buildings carried out either by Maio et al. (2015) or Simões et al. (2015), for example, were
conducted using the 3Muri ® software (STADATA 2011), which is based on the nonlinear
macro-element model proposed by Gambarotta and Lagomarsino (1996) and further
modified by Penna (2002). In these studies, the performance point or target displacement
of the structure was computed from the intersection of the capacity curve of the structure
and the response spectrum. The capacity curve was obtained by converting the pushover
curve from the original MDoF model to an equivalent SDoF model, using the N2 Method,
originally proposed by Fajfar (1999) and adopted by both Eurocode 8 (CEN 2004) and the
Italian Seismic Code (NTC 2008). Del Gaudio et al. (2015) have instead applied the
Incremental N2 Method (IN2), which provides the relationship between seismic demand and
seismic intensity. The basic difference between the N2 Method and the original capacity
spectrum method (CSM), proposed by Freeman (1998) and widely implemented in the USA
and Japan, is in the inelastic demand definition. While the inelastic demand in the N2
Method is defined by an inelastic spectrum, in the CSM the inelastic demand is defined by
means of an equivalent highly-damped elastic spectrum.
Nevertheless, as the seismic demand is determined for an equivalent SDoF system, these
procedures are naturally based on a single-mode response with time-invariant displacement
shape, and therefore are all subjected to some limitations A comparison of these methods
showed that the N2 Method gave more accurate results than those provided by the CSM,
when considering the results from nonlinear dynamic analysis as the baseline method (Silva
et al. 2013).

3.5. Shear failure


Reinforced concrete and masonry walls do not reach failure only in flexural modes but they
can also fail in shear modes. If the shear strength of a wall is reached before its flexural
strength, failure occurs very suddenly, with little warning, and the ductility of the wall is
limited. Shear failure in members can occur in RC buildings designed without considering
the effect of horizontal actions, or in buildings with low concrete strength or without
sufficient horizontal reinforcement. Even though this failure mode can induce important
effects on fragility curves, hastening the attainment of higher damage levels, only very few
methodologies treat this issue explicitly.

28
3.6. Out-of-plane response of masonry buildings
The majority of the existing methodologies and software currently available for seismic
assessment of masonry structures usually considers only in-plane failure modes. However,
according to past observations after major earthquakes, the out-of-plane failure of masonry
walls has been stressed out as one of the most dominant modes of failure for unreinforced
masonry buildings (Ceran 2010, Ferreira 2015). The structural deficiencies that lead to this
type of failure are quite a few, from the poor wall-to-floor and wall-to-wall connections due
to insufficient anchorage, the absence of horizontal rigid diaphragms resulting in high wall
slenderness ratio, to poor material quality (Ceran 2010). Hence it is misleading to evaluate
the seismic vulnerability of masonry structures without considering their out-of-plane
behaviour. In addition to this, out-of-plane wall failures impose a significant risk to the
people living in these buildings, since they may be trapped by (parts of) overturned walls
(Ceran 2010).
The reason for which out-of-plane response of unreinforced masonry walls is often
disregarded is that, as stated by Paulay and Priestley (1992), it is one of the most complex
and ill-conditioned problems in seismic analysis of building structures, being this behaviour
governed both by strength and stability (Ceran 2010). Hence, many different approaches
for the analysis of the out-of-plane seismic behaviour of unreinforced masonry structures
have been developed to solve this problem during the last decades. Some of these
approaches were addressed and reviewed by Ferreira (2015), which grouped them in
numerical, experimental and analytical, as demonstrated in Figure 16, in which
experimental works can also be used for the calibration and validation of both numerical and
analytical methodologies.
A different line of investigation resorting to multibody dynamics theory has been recently
proposed by Costa (2012), for the numerical analysis of the out-of-plane dynamic behaviour
of unreinforced masonry buildings, in which portions of masonry walls considered
representative of the out-of-plane local mechanisms activated by seismic loads, are
modelled as kinematic chains (normally assumed as infinitely rigid bodies) whose nonlinear
behaviour is concentrated at the contact regions. According to Ferreira (2015), the main
advantages on the use of multibody dynamics to simulate complex local mechanisms rely on
the time-efficiency and small number of input parameters required. Moreover, a further
disadvantage of this technique is that it requires the definition of a realistic overturning
mechanism in order to create the multibody model, based on expert judgement or some
other type of analysis (Ferreira 2015).

Figure 16. Existing techniques for assessing the out-of-plane behaviour of unreinforced masonry
structures (Ferreira 2015).

29
In the previously referred SP-BELA method, for example, the out-of-plane failure
mechanism is defined only considering the boundary conditions of walls supported at the
base and free at the top, and walls supported at both sides with or without axial load (Borzi
et al. 2008c). However, complex out-of-plane failure mechanism such as those described in
Restrepo-Velez and Magenes (2004) were not included as a consequence of the limited
information that is generally available at an urban scale.
In literature most approaches refer to equivalent frame systems to definition of the in-plane
behaviour of walls (Erberick 2008, Lagomarsino et al. 2006, Barbat 2004). Hence, the
resulting fragility curves ignore the possibility of out-of-plane behaviour, which is
particularly important in the case of URM buildings. The studies carried out by Tomaževic et
al. (1991), D’Ayala et al. (1997), Borzi et al. (2008), Karantoni et al. (2014), outlined the
relevance of out-of-plane failure on the overall fragility of masonry structures, as for
instance in the case of buildings with flexible roofs. Recently an increasing number of
authors have started considering simple out-of-plane modelling with limit-state mechanical
approaches. However, the quality of the connections among walls is often overlooked and
the corresponding fragility overestimated (D’Ayala and Meslem 2012).

3.7. Horizontal diaphragms in masonry buildings


The experience gathered from past earthquakes has proved that, among other parameters
that influence the resistance of existing masonry buildings, the effectiveness of the
connections between the horizontal (floors) and vertical (walls) elements together with the
in-plane stiffness of floors greatly affects the seismic response of the structure (Rota et al.
2011). Although generally conceived to carry vertical loads, the horizontal elements play a
significant role not only in the distribution of horizontal inertia forces during an earthquake
onto the walls proportional to their stiffness, but also in the prevention of out-of-plane
mechanisms enabling a box-type of behaviour of the structure (Tomaževic et al. 1991). In
case of poor wall-to-floor connection, separation of the elements and overturning of walls
orthogonal to the seismic motion may occur, even at low to moderate intensity of shaking
Senaldi et al. (2014). Therefore, it is important to understand how the variability of stiffness
of different diaphragm typologies influences the global seismic response of the structure and
to account for this variability when developing fragility curves for masonry buildings.
Ceran (2010) has generated fragility curves for assessing the out-of-plane behaviour of
masonry buildings, also considered different sets of curves to account for both the presence
of rigid RC slabs and wooden flexible diaphragms. The meaningful influence of the type of
wall-to floor connection, slenderness ratio of the wall and the strength of masonry wall, over
the damage state probabilities was one of the main conclusions from this work. Moreover,
results have proved that non-engineered masonry buildings with very poor wall-to-floor
connections, in which the walls imitate cantilever-like behaviour, are extremely vulnerable
to even low- to-moderate levels of seismic action. Thus, as the out-of-plane behaviour may
be more critical than the in-plane behaviour, it should be considered when deriving fragility
curves for URM buildings.

3.8. Geometrical irregularities


In-height and in-plan irregularities, both in terms of stiffness and mass are capable to
generate stress distributions significantly different from those expected in regular
structures, which might cause severe damage or even the collapse of the structure.

30
Empirical observations of structures damaged by recent earthquakes have demonstrated
that generally simple and regular structural systems present a better seismic behaviour,
while complex and irregular systems lead to a poor behaviour. In addition, the absence of
masonry infill walls at the ground floor level is a very common practice in RC buildings due
to architectural and commercial purposes. However, this is responsible for triggering soft-
storey collapse mechanisms (e.g. Verderame et al. 2011; De Luca et al. 2014). In the
studies carried out by Erberik (2008), Oropeza et al. (2010) and Polese et al. (2008) in plan
irregularities were taken into account when generating fragility curves for URM buildings.
Rota et al. (2014) in turn, considered both in plan and in height irregularities. These studies
have confirmed the higher fragility of irregular structures, both in plan and in height.

3.9. Seismic demand and site-specific records


The seismic demand for a specific site is usually determined by means of code-based
spectra or real ground motion records. The assessment of the structural response through
dynamic analysis requires seismic input that should reflect the hazard as well as the near-
surface geology at the site. Generally, the signals that can be used for the seismic structural
analysis are of three types: artificial waveforms (compatible with a code response
spectrum); simulated accelerograms (obtained via modelling of the seismological part and
possibly accounting for the path and site effects), and natural records.
Like many codes worldwide, Eurocode 8 (CEN 2004) allows the use of real ground motion
records for the seismic assessment of structures. The main condition to be satisfied by the
chosen set is that the average elastic spectrum does not underestimate the code spectrum,
with a 10% tolerance, in a broad range of periods depending on the structure's dynamic
properties. In particular, Eurocode 8 allows employment of all three types of accelerograms
listed above as an input for seismic structural analysis. If at least seven records are used,
the mean response is considered, while for three to six records, the maximum response
should be considered. Appropriately selected real accelerograms are becoming the most
attractive option to get unbiased estimations of the seismic demand (Iervolino et al. 2008).
When assessing sites where no recorded accelerograms are available, it is common to
consult ground motion databases, such as the European strong-motion database
(www.isesd.hi.is) and the PEER Ground-Motion Database, for accelerograms that best match
the conditions of the site, in terms of magnitude, M w, rupture distance, rrup, and average
shear-wave velocity between 0 and 30 m, vs,30. These databases provide free access to
thousands of strong-motion records from shallow crustal earthquakes and also easy-to-use
online tools for the selection of records according to criteria related to earthquake scenario,
local site conditions and response spectra. Even though using site-specific ground motions is
always preferable in order to reduce the uncertainty in demand of the analysis, one might
take advantage of collected accelerograms from these databases only when they present
similar characteristics as the assessment site, which is not always observed. Moreover, the
selection of suitable accelerograms requires a certain level of experience on both geological
and seismic hazard context of the specific site under assessment.

3.10. Definition of EDP values at damage thresholds


When detailed numerical models of the building are used, the damage in each structural
element is obtained through static or dynamic nonlinear analysis, in which damage state
attribution could be made as a function, for instance, of the its displacement capacity when
subjected to horizontal excitation (Lagomarsino and Cattari 2014). However, it is worth
noting that the identification of discrete damage states is not an easy task.

31
Engineering Demand Parameters are measures of the structural response that can be
obtained from the results of structural analyses. On one hand, typical choices from the
examined methodologies that have performed Nonlinear Dynamic Analysis are the Inter-
storey Drift (ISD), Global Drift (GD) or the Park and Ang (1985) damage index, among
others. On the other hand, top displacement is the most commonly utilized EDP from
Nonlinear Static (Pushover) Analysis. Threshold values of EPDs can be estimated either for
each individual building, or for building typologies.
The nonlinear analysis usually provides specific values of EDPs, normally at the level of
elements, which describe changes in structural behaviour. It is hence common practice to
use these EDP values or a combination of them to also define the threshold for each
damage state at local or global level. The definition of capacities for each building provides
the flexibility of tailoring the damage threshold values to a given building or sub-class of
building typologies and introduces capacity-demand correlations that may have a major
influence in the fragility analysis results.
Alternatively, one might employ the proposed simplified formula and relations from
literature, to directly estimate global damage state thresholds and the corresponding
median capacity values. Relations commonly used are expressed as functions of yield and
ultimate roof displacement.
EDP limit values may also be estimated from a (simplified bilinear) capacity curve of a
building. For example, Lagomarsino and Giovinazzi (2006) identified four damage states on
the global capacity curve and associated to each a median value of spectral displacement.
Moreover, Cattari and Lagomarsino (2012) have proposed a multi-scale approach for
masonry buildings that defines limit states thresholds also based on the capacity curve, by
checking the spread of damage in masonry elements (piers and spandrels), the inter-storey
drift in masonry walls, and the global behaviour of the building (described by its capacity
curve).
Likewise, D’Ayala (2013) has suggested four damage state thresholds for the generation of
fragility curves for masonry buildings, this time obtained through laboratory or in-situ
experimental tests reported in literature on whole buildings or full-size walls with different
failure modes. These values were used when developing the fragility curves with the FaMIVE
procedure.
Regarding the evaluation of different limit states at the level of the structure there is still a
lack of clear guidance in the literature.

3.11. Form of relation


In the vast majority of the existing fragility models and related methodologies, a lognormal
function defined by a mean and standard deviation is employed to represent the conditional
probability of exceeding a damage state, as in the case of the fragility curves built by Barbat
et al. (2006), presented in Figure 17 (left). Also in the FAMIVE method, fragility curves for
different limit states are obtained by computing the median and standard deviation values
of the performance point displacements for each building and by deriving lognormal
distributions (D’Ayala et al. 2014). However, as demonstrated by Ceran (2010), this
probability can also be obtained through a standard normal cumulative distribution function
for the construction of fragility curves, which in the case of Figure 17 (centre), are referring
to the ultimate limit state of the critical walls of three different Turkish case study masonry
buildings. Additionally, this probability can be obtained through a set of discrete values, as
observed in D’Ayala et al. (1997), which have generated fragility curves for the Alfama
District of Lisbon, illustrated in Figure 17 (right).

32
Figure 17. Continuous fragility curves for lognormal (left) and normal (centre) functions, and discrete
fragility curves (right), developed by Barbat et al. (2006), Ceran (2010) and D’Ayala et al. (1997),
respectively.

3.12. Intensity Measure


In order to understand the role that intensity measure (IM) plays on the seismic risk
assessment process, it is important to recall the definition of fragility curve, which
represents the conditional probability of reaching a given damage state for a given hazard
level characterized by a determined IM. Therefore, IMs are a key point on the derivation
process of fragility curves.
Although IMs can be chosen in different ways, for the analysis to be robust, they are
required to accurately represent the relevant seismological properties of ground motion
related to site and source (Luco and Cornell 2007). Given the spatial distribution of the
building stock, it is practical and advisable to use IMs for which ground motion prediction
equations (GMPEs, also known as attenuation relationships) are available. The selection of
IMs should aim at a reduced record-to-record variability of the selected EPD, enabling thus
its evaluation with a small number of time-history analyses (D’Ayala et al. 2014). Thus, the
identification of proper IM is subject to different constraints, which are related to the
adopted hazard model, to the typology of the exposed asset but also to the availability of
data for all different exposed assets (Lagomarsino and Cattari 2014).
As one can observe from the reviewed methodologies in Table 1, at the end of this chapter,
the most recurrently used IMs are the peak ground acceleration, peak ground velocity, peak
ground displacement, and (pseudo) spectral acceleration values. However, according to
Gehl et al. (2012), the standard method to develop fragility curves neglects the uncertainty
in the estimated damage caused by the use of a single IM, which cannot represent all
characteristics of the ground motion. In this sense, some efforts have been made recently
to consider the effect of several IMs on structural damage (Baker and Cornell 2005; Kafali
and Grigoriu 2007; Seyedi et al. 2010; Koutsourelakis 2010).

3.13. Sample size


The sample size accounts for the number of buildings (population) under analysis. It is
extremely difficult to generalise results or establish comparisons based on methodologies in
which only an individual building (or case study) is analysed. Although this generalisation
and comparison of results is always questionable, it is clear that the larger the population of
buildings analysed the higher the chances for generalising the respective results.
For the selection of the sample, it is important to identify the main parameters that
influence the seismic response of a building class, such as the lateral load resisting system,

33
mechanical properties of materials, geometric configuration, dimensions and detailing of
elements (D’Ayala et al. 2014). These parameters may show considerable scatter. Thus, a
comprehensive knowledge of the probabilistic distribution of the most important structural
parameters defining the population of buildings in the class of interest is needed. Based on
the fragility curves reviewed in this report, the most important characteristics that typically
are accounted for include:
• Building height, in discrete values or range of storeys;
• Design base shear and deformation capacity of the building;
• In plan irregularity;
• In height irregularity, defined for example by the presence of a soft/weak storey and
the ratio of its strength or stiffness to that of the adjacent storey;
• Material properties.
Monte Carlo simulations typically involve the generation of a large sample, which can
accurately represent the underlying population. This type of simulation it is not yet practical,
mainly because of the effort involved in the criterion, modelling and analysis of the sample,
unless automated software and generous computing resources are available. Thus, Monte
Carlo simulations are more practical for the simplest SDoF models. Its application together
with improved sampling strategies for example Latin Hypercube Sampling (LHS), rather
than classic random sampling, drastically reduces the computational cost and offers insights
into the actual population statistics. Lastly, it is noted that appropriate skills and expertise
are necessary to perform Monte Carlo simulations.
At this stage, the main features associated with the methodologies considered in this report
were approached and described. Thus, in the next Chapter 4, the criteria defined for the
qualitative evaluation of those fragility curves will be presented.

3.14. Summary table of the reviewed fragility curves


The following Table 1 displays an informative summary of some of the most important
features of each fragility curve methodology analysed in this report. Briefly, these are
grouped in four main categories, each containing subcategories: general features; capacity;
demand, and fragility curves’ inherent characteristics. In the following paragraphs the
authors will describe the taxonomy herein utilised in Table 1, even though most of these
features were previously addressed throughout the current Chapter 3.
The first category, termed general features, includes the respective reference of each
methodology, its geographical applicability, class of structures analysed and structural
system typology. In what regards the class of structures, EB stands for existing buildings
whose dimensions, materials and properties are known and used in the analyses, and PR
stands for prototype buildings. The structural system typologies analysed in this work
included unreinforced masonry (URM) and reinforced concrete (RC) structures. As already
mentioned, RC buildings include the following subcategories: moment-resisting frames; dual
frame-wall systems; infilled frames, and precast systems.
The second category comprises the following features related to capacity: components for
response analysis; number of floors; analysis type; model type; shear failure; out-of-plane
mechanism; horizontal diaphragm; geometrical irregularities, and engineering demand
parameter (EDP). In what concerns the components considered for response analysis, S and
NS stand for structural and non-structural elements, respectively. In relation to the different
type of analyses found in the reviewed methodologies, the following nomenclature was
used: NLD for nonlinear dynamic analysis; NLS for nonlinear static analysis, SMM-NLS for
simplified mechanic methods, and LS for linear static analysis. In what regards the model

34
type chosen for the analysis, three different designations were considered: MDoF (multi-
degree of freedom system); reduced MDoF, and SDoF (single-degree of freedom system)
models. For the shear failure, out-of-plane mechanism, horizontal diaphragms and
geometrical irregularities subcategories, checkmarks  and  were used to identify if these
features were taken into consideration in the reviewed methodologies. Moreover, the
symbol (-) was used when a particular subcategory is not applicable to a certain
methodology. Finally, in what concerns the engineering demand parameters (EDP), the
following nomenclature was used: RD (roof displacement); TD (top displacement); ISD
(inter-storey drift); GD (global drift); VLS,i (median base shear capacity at the i th limit
state); VD (median base shear demand given the PGA); MPR (maximum plastic end rotation
of beams), and MFA (maximum floor acceleration).
The third category is related to the seismic demand-related features and includes both
seismic demand and site-specific subcategories. As the reviewed methodologies use either
ground motion records, including both real and synthetic records, or code-compliant
spectra, these two broad designations were herein considered. As previously explained, the
site-specific subcategory assesses if the seismic input considered for analysis in a given
methodology is related or not with the specific site in which the buildings under assessment
are located. Thus, checkmarks  and  were used also here to identify if seismic input is
related or nor to the specific site.
The last category includes some of the most important features for deriving fragility curves,
including the damage state thresholds definition, number of damage states, form of
relationship, intensity measure and sample size. As described before in this Chapter 3,
damage state thresholds can be defined by means of either pre-set or custom definitions.
With respect to the form of relationship for the construction of fragility curves, three
different options are usually considered: CL (continuous lognormal); CN (continuous
normal), and DD (discrete distribution). In what regards the intensity measure (IM), the
following taxonomy was used: Sd(T) (spectral displacement); PGA (peak ground
acceleration); PGV (peak ground velocity); PGD (peak ground displacement); CAV
(cumulative absolute velocity); Sa(T) (spectral acceleration); MMI (modified Mercalli
intensity); Sv(T) (spectral velocity), and AI (Arias intensity). Finally, for the size of the
sample considered in each reviewed methodology, the authors attributed the following
classification: several buildings; few buildings, and one single building.

35
Table 1. Main features of the considered methodologies for deriving analytical fragility curves.

General Capacity Demand Fragility curve

Analysis type
Geographical

irregularities
Shear failure

Out-of-plane
for response
Components
applicability

Geometrical

Site specific

relationship

Sample size
mechanism
Model type

diaphragm

thresholds
Number of

Number of
Horizontal
Reference

Structural

Intensity
measure
Damage
analysis

demand

damage

Form of
Seismic
storeys
system

states
states
Class

(IM)
EDP
TR, IT,
Ahmad et al. RC moment- Ground motion Sd(T) and
GR and PR S 2,5,8 NLS SDoF  (-) (-)  TD  Pre-set 5 CL Several
(2010) resisting frame records PGA
SI

Akansel et al. RC dual frame-wall Ground motion PGA, PGV,


TR PR S 3 NLD MDoF  (-) (-)  ISD  Custom 3 CL One
(2012) system records PGD and CAV

Akkar et al. NLD + Ground motion


TR EB RC infilled frames S+NS 2,3,4,5 SDoF  (-) (-)  GD  Custom 4 CL PGV Several
(2005) NLS records

Barbat et al. 2,5,8 + Ground motion


ES EB RC+URM S NLS SDoF  (-)+ (-)+  TD  Pre-set 5 CL Sd(T) Several
(2006) 2,4,6 records

RC moment-
Ground motion
Bilgin (2013) TR PR resisting + dual S 3,4,5 NLS SDoF  (-) (-)  ISD  Custom 4 CL PGV Several
records
frame-wall systems

Borzi et al. RC moment- SMM- Chord Code-based


IT PR S 2,4,8 SDoF  (-) (-)   Custom 4 CL PGA Several
(2007) resisting frame NLS rotations spectra

RC moment-
Borzi et al. SMM- Chord Code-based
IT PR resisting + dual S+NS 4 SDoF  (-) (-)   Custom 4 CL PGA Several
(2008a) NLS rotations spectra
frame-wall systems

Borzi et al. RC moment- SMM- Chord Code-based


IT PR S 2,3,4,5,6,8 SDoF  (-) (-)   Custom 5 CL PGA Several
(2008b) resisting frame NLS rotations spectra

Borzi et al. SMM- Chord Code-based


IT PR URM S 2,3,4,5 SDoF      Custom 5 CL PGA Several
(2008c) NLS rotations spectra

Casotto et al. NLD + MDoF + Maximum top- Ground motion


IT PR RC (precast) S 1  (-) (-)   Custom 5 CL Sa(T) Several
(2015) NLS SDoF drift records

SMM- Base + flexural Code-based


Ceran (2010) TR PR URM S Multiple SDoF      Custom 3 CN PGA Several
NLS strength spectra

D'Ayala SMM- Collapse load Code-based


TR EB URM S Multiple SDoF      Custom 6 CL Sd(T) Several
(2005) NLS factor spectra

Del Gaudio et SMM- Code-based


IT EB RC infilled frames S+NS Multiple SDoF  (-) (-)  TD  Pre-set 6 CL PGA Several
al. (2015) NLS spectra

36
General Capacity Demand Fragility curve

Analysis type
Geographical

irregularities
Shear failure

Out-of-plane
for response
Components
applicability

Geometrical

Site specific

relationship

Sample size
mechanism
Model type

diaphragm

thresholds
Number of

Number of
Horizontal
Reference

Structural

Intensity
measure
Damage
analysis

demand

damage

Form of
Seismic
storeys
system

states
states
Class

(IM)
EDP
RC moment-
Dumova-
resisting + dual Reduced Park and Ang Ground motion
Jovanoska MK EB S 6,16 NLD  (-) (-)   Custom 5 DD MMI Several
frame-wall MDoF damage index records
(2000)
systems

Erberik NLD + Ground motion


TR EB URM S 1,2,3,4,5 SDoF     VLS,i, VD  Custom 3 CN PGA Several
(2008) NLS records

RC moment-
resisting + dual
Fardis et al.  Chord rotation, Code-based
Europe PR frame-wall S+NS 2, 5, 8 NLS SDOF  (-) (-)  Custom 2 CL PGA One
(2012) shear force spectra
systems with
infills
Sa(T), Sv(T),
Hancilar and RC moment- ISD, MPR and Ground motion Sd(T), PGA,
TR EB S 5,10,15,20 NLD MDoF  (-) (-)   Custom 13,11,15 CL Several
Çaktr (2015) resisting frame MFA records PGV, PGD, AI
and CAV

Kappos and RC moment- Low, mid Ratio of roof


Reduced Ground motion
Panagopoulos GR EB resisting + dual- S and high NLS  (-) (-)  displacement  Pre-set 6 CL PGA Several
MDoF records
(2009) frame systems rise to total height

RC dual frame-
Kappos et al. 1-3,4-7,8- Reduced Ground motion
GR EB wall + infilled S+NS NLS  (-) (-)  ISD  Pre-set 6 CL PGA Several
(2003) 19 MDoF records
frame systems

RC dual frame- Reduced


Kappos et al. NLD + Ground motion Sd(T) and
GR EB wall + infilled S+NS 4-7,>8 MDoF +  (-) (-)  ISD  Pre-set 6 CL Several
(2006) NLS records Sa(T)
frame systems SDoF
Reduced
Kappos et al. NLD + Ground motion
GR EB URM S Multiple MDoF +     RD, ISD, TD  Pre-set 5 CL Sd(T) Several
(2007) NLS records
SDoF

Karantoni et Extent of Code-based


GR PR URM S Multiple LS MDoF      Custom 5 CL Sa(T) Several
al. (2014) damage spectra

Kircil and RC moment- Reduced Ground motion


TR EB S 3,4,5,6,7 NLD  (-) (-)  ISD  Custom 2 CL Sa(T) Several
Polat (2006) resisting frame MDoF records

Lagomarsino Low, mid


SMM- Code-based
and Cattari IT PR URM S and high SDoF     ISD  Custom 4 CL PGA Few
NLS spectra
(2014) rise

SMM- Code-based
Lang (2002) CH EB RC+ URM S 2-7 SDoF  (-)+ (-)+  TD  Custom 5 CN Sd(T) Several
NLS spectra

Oropeza et Code-based
CH EB URM S 15 NLS SDoF     TD  Pre-set 5,6 CL Sd(T) One
al. (2010) spectra

37
General Capacity Demand Fragility curve

Analysis type
Geographical

irregularities
Shear failure

Out-of-plane
for response
Components
applicability

Geometrical

Site specific

relationship

Sample size
mechanism
Model type

diaphragm

thresholds
Number of

Number of
Horizontal
Reference

Structural

Intensity
measure
Damage
analysis

demand

damage

Form of
Seismic
storeys
system

states
states
Class

(IM)
EDP
Özer and
RC moment- Reduced Ground motion
Erberik TR PR S 3,5,7,9 NLD  (-) (-)  ISD  Custom 4 CL PGV Few
resisting frame MDoF records
(2008)

Pagnini et al. SMM- Code-based


PT EB URM S 4 SDoF     Ultimate drift  Custom 5 CL Sd(T) One
(2011) NLS spectra

Pasticier et NLS + Reduced Ground motion


IT EB URM S 3     TD  Custom 4 CL PGA One
al. (2008) NLD MDoF records

Pitilakis et al. RC moment- Ground motion


GR EB S 3,4,9 NLD MDoF  (-) (-)  ISD  Custom 3 CL PGA Few
(2014) resisting frame records

Polese et al. RC moment- Code-based


IT EB S 1-3,4-6,>7 NLS SDoF  (-) (-)  TD  Pre-set 5 CL Sd(T) Few
(2008) resisting frame spectra

Rota et al. NLD + Ground motion


IT PR URM S 3 SDoF     GD  Custom 5 CL PGA One
(2010) NLS records

Rota et al. Code-based


IT PR URM S 2-4 NLS SDoF     TD  Custom 4 CL PGA Few
(2014) spectra

Reduced
Silva et al. RC moment- NLD + Ground motion
TR EB S 4 MDoF+  (-) (-)  TD  Custom 4 CL Sa(T) Few
(2013) resisting frame NLS records
SDoF
Reduced
Silva et al. 1,2,3,4,5- NLD + Ground motion
PT EB RC+URM S MDoF +  (-)+ (-)+  GD and ISD  Pre-set 5 CL Sa(T) Few
(2014) 7,>8 NLS records
SDoF

Simões et al. Code-based


PT EB URM S 4,5 NLS SDoF     TD  Custom 5 CL Sa(T) Few
(2015) spectra

RC moment-
resisting + dual
Tsionis and  Chord rotation, Code-based
GR PR frame-wall S+NS 2, 5, 8 NLS SDOF  (-) (-)  Custom 2 CL PGA One
Fardis (2014) shear force spectra
systems with
infills
RC moment-
Park and Ang
Vacareanu et resisting frame + Code-based
RO EB S+NS 8,11,12,13 NLS SDoF  (-) (-)  damage index,  Pre-set 5 CL Sd(T) Few
al. (2004) dual frame-wall spectra
ISD
systems
RC moment-
Vacareanu et resisting frame + Code-based
RO EB S+NS 11 NLS SDoF  (-) (-)  TD  Pre-set 5 CL Sd(T) One
al. (2007) dual frame-wall spectra
systems

38
4. Criteria and evaluation of fragility curves

Having identified and discussed the most important features influencing the accuracy,
complexity and uncertainty associated with the methodologies used to derive fragility
curves, in this Chapter the criteria for the qualitative evaluation of these methodologies,
will be described and applied. The proposed criteria are based on the work carried out by
Rossetto et al. (2014), from which we adopt similar principles to cover four fundamental
categories related to capacity, demand, methodology for fragility analysis, and
uncertainty, as previously discussed in Chapter 3. However, the criteria in the present
report differ in some aspects. In contrast to Rossetto et al. (2014), which have
developed a quantitative (scoring) method for combining fragility curves, the aim of the
present report is to provide a set of qualitative criteria to support the selection process
of the most appropriate fragility curves for the European building stock, which was
intended to be exclusively qualitative, and not quantitative.
The following criteria were examined for each of the above-mentioned categories,
presented in Table 1 of the previous Chapter 3. To each set of fragility curves, low (L),
medium (M) or high (H) rating is assigned to each criterion, as shown in Table 2 and
further detailed in the following sections.

Table 2. Qualitative assessment criteria and respective rating.

Rating
Category Evaluation criteria
Low (L) Medium (M) High (H)
Non-structural elements No (-) Yes
Classes of building height 1 2, 3 >3
Analysis type NLS-SMM, LS NLS NLD
SDoF Reduced MDoF
Model type
Capacity models MDoF models models
Shear failure No (-) Yes
Out-of-plane mechanism (URM) No (-) Yes
Horizontal diaphragms (URM) No (-) Yes
Geometric irregularities No (-) Yes
Code-based <7 ≥7
Seismic demand
Demand spectra accelerograms accelerograms
Site-specific No (-) Yes
Damage state thresholds definition Pre-set (-) Custom
Intensity measure 1 2,3 >3
One Few Several
Methodology Sample size
building buildings buildings
One-index Three-index Multiple-index
Sampling method
building building building
Capacity No (-) Yes
Uncertainty Seismic demand No (-) Yes
Damage state thresholds No (-) Yes

It is not the purpose of this report to rank and identify the ‘best’ set of fragility curves,
as the authors acknowledge the fact that all methodologies present advantages and
drawbacks. Moreover one cannot compare them globally, because each methodology
was developed for very particular conditions, such as, for instance, the assessment
scale, the analysis computational effort, human and time resources consumption.
Nevertheless, according to Calvi et al. (2006) an optimal or ideal fragility assessment
methodology should feature the following main characteristics:

39
• The most recent developments in the field of seismic hazard assessment should
be incorporated;
• All sources of uncertainty should be explicitly accounted for;
• The methodology should be easily adaptable to the different construction
practices around the world, as well as allow for the inclusion of new construction
types and the influence of retrofitting on the response of existing structures;
• A balance should be struck between the computational intensity and the amount
of detailed data that is required and the consequent degree of confidence in the
results.

It is unlikely that a single methodology can fulfil all of these requirements and therefore,
reliable fragility assessment of a given region is likely to request the employment of at
least two different approaches, which should complement and/or verify each other (Calvi
et al. 2006).
The criteria herein proposed aim to provide guidance on the selection of fragility curves
for the European building stock, based on a two-step selection procedure. Firstly, one
should short down the spectra of available fragility curves by considering those that
apply to the geographical region and building typology of interest. Secondly, one should
select the fragility curves that better fit the specific purpose, based on the four
fundamental categories that constitute the proposed evaluation criteria. In the following
sections, detailed information will be given concerning the assessment of each single
feature.

4.1. Capacity
As mentioned in the previous Chapter 3, the consideration or not of non-structural
components in fragility assessment of buildings is generally more relevant in the case of
RC buildings. Here, the consideration of the contribution of masonry infill walls in the
analysis model, taken as non-structural components, has been highly recommended due
to its significant influence on the lateral resistance of a reinforced concrete frame.
Therefore, according to this criterion, the reviewed fragility curves that have considered
the presence of non-structural components in their analyses when assessing RC
buildings were given a high rating (H), while those that have only considered the
structural components were assigned a low (L) rating.
The criterion defined for the number of storeys classifies the reviewed fragility curves
according to the variability of the number of storeys considered in the analysis. This is
relevant to understand the different behaviour of low-, mid- and high-rise buildings,
which naturally have distinct fragilities. Thus, fragility curves in which the variability in
terms of number of storeys was not considered in the analyses were given a low rating
(L). Accordingly, fragility curves that have considered up to three different numbers of
storeys in their analyses were given a medium rating (M). Finally, fragility curves that
have considered in their analyses more than three different numbers of storeys were
given a high rating (H).
Considering that the reliability associated to the type of analysis chosen for structural
assessment is strongly related to the level of detail of that analysis, in the present
criteria the authors assigned NLD analysis high (H), NLS analysis medium (M) and NLS-
SMM or LS analyses low rating (L).
Based on the same reasoning as for the analysis type, the model types selected for
structural assessment were evaluated as a function of their level of detail. Thus, MDoF
models were assigned high (H), reduced MDoF models medium (M), and SDoF models
low rating (L).
As mentioned in Chapter 3, shear failure in members can occur in buildings that are
designed for horizontal actions or in buildings with low shear resistance. This failure
mode can have important effect on fragility curves as it hastens the attainment of higher

40
damage levels. Therefore, the fragility curves in which shear failure mode was
considered in the analyses were assigned high rating (H). In contrary, those fragility
curves in which this failure mode was not considered in the analyses or no clear
explanation is provided about its consideration were assigned a low rating (L).
Acknowledging that the out-of-plane response of URM buildings is often neglected in
numerical modelling, despite representing one of the dominant modes of failure, the
fragility curves that accounted or not for this particular failure mode in the analyses were
given high (H) and low rating (L), respectively.
Given that the stiffness of different diaphragm typologies influences the global seismic
response of URM buildings, the criterion defined to evaluate the reviewed fragility curves
was based on whether the variability of the horizontal diaphragms was considered or not
in the analysis, assigning high (H) and low rating (L), respectively.
Finally, as mentioned before in this report, irregularities are capable to generate stress
distributions significantly different from those expected in regular structures, which
might cause severe damage or even the collapse of the structure. However, the effect of
these irregularities is not always considered when performing seismic fragility
assessment of both RC and URM buildings. Hence, high (H) and low rating (L) was
assigned to the reviewed methodologies that have considered or not geometrical
irregularities, respectively.

4.2. Demand
The criterion established for the description of the seismic demand was based on the
Eurocode 8 recommendations (CEN 2004). Thus, a high rating (H) was assigned to
methodologies in which at least seven real ground motion records were used to estimate
the seismic demand. Then, methodologies which have determined the seismic demand
for a specific site by means of considering less than seven real ground motion records
were assigned as medium (M). Finally, methodologies where code-based spectra were
considered were evaluated as low (L).
In what concerns the criterion for the site-specific input, the reviewed methodologies are
given high rating (H) when the seismic demand is related to the specific site of the
assessed structures and low (L) in the contrary case, even when covering similar
magnitudes, rupture distances and average shear-wave velocities.

4.3. Fragility curve (methodology)


Bearing in mind the two approaches for the definition of EDPs threshold values
mentioned and described in the previous Chapter 3, the calculation of custom values for
each specific building was assigned as high (H), while the use of pre-set values for
building typologies was given low rating (L).
Regarding the IM used for the derivation of fragility curves, low (L), medium (M) and
high (H) rating was assigned those that use respectively one, two and three or more
IMs. This criterion does not address the issue of correlation of damage to the IM and the
effort needed to calculate the IM, which are difficult to assess even qualitatively and are
often conflicting (e.g. spectral values and composite IMs are better correlated to
structural damage than peak ground acceleration, but require additional calculations).
As it is found extremely difficult to generalise results or establish comparisons based on
methodologies in which only an individual building (or case study) was analysed, the
evaluation of this subcategory aims at distinguishing from the reviewed methodologies
those that are carried out for larger samples, increasing its applicability and
generalisation to particular building typologies of the European building stock. Hence,
samples sizes of several (hundreds) of buildings were assigned as high (H), while those

41
with few (dozens) of buildings were assigned as (M). Finally, those methodologies in
which only one single building was analysed were evaluated as low (L) in this
subcategory.

4.4. Treatment of uncertainty


Uncertainty in capacity is introduced either through geometrical, mechanical, structural
or modelling parameters. Usually, in order to create a number of building models for the
structural analysis, uncertainty in capacity is addressed by treating both geometric and
material properties, mass and damping, as random variables. The treatment of
uncertainty in capacity was evaluated as sufficient and marked high (H) if at least one of
the following sources were accounted for the derivation of fragility curves: mechanical
properties; geometric parameters; structural detailing, or numerical modelling. On the
contrary, if none of the previous sources of uncertainty in capacity were considered, the
rating was evaluated as low (L).
The uncertainty in the demand is associated with the natural variability of ground motion
characteristics, which comprises the variability of the seismic source, path attenuation
and site effects of the seismic event. This type of uncertainty is typically taken into
account either through the selection (and scaling) of natural ground motion records or
by generating artificial records. The treatment of uncertainty in demand was herein
evaluated as high (H) if record-to-record variability is considered.
Uncertainty in the definition of damage state thresholds is often neglected in fragility
assessment procedures found in the literature. However, in a few cases a probability
distribution is assigned to the damage state thresholds. In the present criterion,
methodologies in which uncertainty in damage state thresholds was considered were
assigned as high rating (H) and those that haven’t considered it in were given low rating
(L).

42
Table 3. Application of the qualitative assessment criteria on the reviewed fragility curves.

General Capacity Demand Fragility curve Uncertainty

Analysis type
Geographical

irregularities
Shear failure

Out-of-plane
for response
Components
applicability

Geometrical

Site specific

Sample size
mechanism
Model type

diaphragm

thresholds

thresholds
Horizontal
Reference

Structural

Classes of

Intensity
measure

Capacity
Damage

Damage
analysis

building

demand

demand
Seismic

Seismic
system

height

states
Class

(IM)
TR, IT,
Ahmad et al.
GR and PR RC L M M L L (-) (-) H H L L M H H H H
(2010)
SI

Akansel et al.
TR PR RC L L H H L (-) (-) H H H H H L H L L
(2012)

Akkar et al.
TR EB RC H H H L L (-) (-) L H H H L H H H H
(2005)

Barbat et al.
ES EB RC+URM L+H M M L H (-)+L (-)+H H H L L L H L H H
(2006)

Bilgin (2013) TR EB RC L M M L H (-) (-) L H H H L H H H H

Borzi et al. (2007) IT PR RC L M L L H (-) (-) L L H H L H H H H

Borzi et al.
IT PR RC H L L L L (-) (-) L L H H L H H H H
(2008a)

Borzi et al.
IT PR RC L H L L H (-) (-) L L H H L H H H H
(2008b)

Borzi et al.
IT PR URM H H L L H H H L L H H L H H H H
(2008c)

Casotto et al.
IT PR RC L L H H L (-) (-) L H L H L H H H H
(2015)

Ceran (2010) TR PR URM H H L L H H H H L H H L H H H H

D'Ayala (2005) TR EB URM H H L L L H L L L H H L H H H L

Del Gaudio et al.


IT EB RC H H L L L (-) (-) L L H H L H H H L
(2015)

43
General Capacity Demand Fragility curve Uncertainty

Analysis type
Geographical

irregularities
Shear failure

Out-of-plane
for response
Components
applicability

Geometrical

Site specific

Sample size
mechanism
Model type

diaphragm

thresholds

thresholds
Horizontal
Reference

Structural

Classes of

Intensity
measure

Capacity
Damage

Damage
analysis

building

demand

demand
Seismic

Seismic
system

height

states
Class

(IM)
Dumova-
MK EB RC L L H M L (-) (-) L H H H L H L H L
Jovanoska (2000)

Erberik (2008) TR EB URM H H H L H L L H H H H L H H H H

Fardis et al.
Europe PR RC H M M L H (-) (-) H L L H L L H H L
(2012)

Hancilar and Çaktr


TR EB RC L H H H L (-) (-) L H H H H H H H H
(2015)

Kappos and
Panagopoulos GR EB RC L M M M L (-) (-) L H H L L H H H H
(2009)

Kappos et al.
GR EB RC H M M M L (-) (-) H H H L L H H H H
(2003)

Kappos et al.
GR EB RC H M H M H (-) (-) L H H L M H H H H
(2006)

Kappos et al.
GR EB URM H H H M L L L L H H L L H L H L
(2007)

Karantoni et al.
GR PR URM H H L H H H H L L L H L H H H H
(2014)

Kircil and Polat


TR EB RC L H H M L (-) (-) L H H H L H L H H
(2006)

Lagomarsino and
IT PR URM H M L L H L L H L H H L M H H H
Cattari (2014)

Lang (2002) CH EB RC+URM L+H H L L H (-)+H (-)+L L L H H L H L H H

Oropeza et al.
CH EB URM H L M L L L L H L H L L L H L L
(2010)

Özer and Erberik


TR PR RC L H H M L (-) (-) L H H H L M H H H
(2008)

44
General Capacity Demand Fragility curve Uncertainty

Analysis type
Geographical

irregularities
Shear failure

Out-of-plane
for response
Components
applicability

Geometrical

Site specific

Sample size
mechanism
Model type

diaphragm

thresholds

thresholds
Horizontal
Reference

Structural

Classes of

Intensity
measure

Capacity
Damage

Damage
analysis

building

demand

demand
Seismic

Seismic
system

height

states
Class

(IM)
Pagnini et al.
PT EB URM H L L L L L L L L H H L L H L L
(2011)

Pasticier et al.
IT EB URM H L H M H L L L H H H L L L L L
(2008)

Pitilakis et al.
GR EB RC L M H H L (-) (-) L H H H L M H H H
(2014)

Polese et al.
IT EB RC L M M L L (-) (-) H L H L L M H H H
(2008)

Rota et al. (2010) IT PR URM H L H L H L H L H H H L L H L H

Rota et al. (2014) IT PR URM H M M L H L L H L H H L M H H H

Silva et al. (2013) TR EB RC L L H M L (-) (-) L H H H L M H H H

Silva et al. (2014) PT EB RC+URM L+H H H M L (-)+L (-)+L H H H L L M H H H

Simões et al.
PT EB URM H M M L H L H H L H H L M H H L
(2015)

Tsionis and Fardis


GR PR RC H M M L H (-) (-) H L L H L L H H L
(2014)

Vacareanu et al.
RO EB RC H H M L L (-) (-) L L H L L M H H H
(2004)

Vacareanu et al.
RO EB RC H L M L L (-) (-) L L H L L L L L L
(2007)

45
46
5. Current trends in seismic fragility curves

This Chapter aims to present the main trends possible to extract from the existing
methodologies for deriving analytical fragility curves for the European building stock,
based on the evaluation criteria described in Chapter 4. Even though the following
output is not exhaustive, this exercise is highly interesting for perceiving how the
scientific community in general is approaching this issue in each of the its main aspects.
It would be indeed of great interest to extend this exercise to all the remaining
approaches used to derive fragility curves, including also other methodologies developed
worldwide.
Figure 18 depicts the distribution of the examined fragility curves in terms of date (year)
of publication. Globally one can observe that the number of studies roughly doubled in
ten years, which somehow evidences that the study of fragility curves is a topic of
growing interest and investment by the scientific community and founding
X institutions.
8

6
Number of studies

0
2000

2001

2002

2003

2004

2005

2006

2007

2008

2009

2010

2011

2012

2013

2014

2015

Year

Figure 18. Distribution of the examined studies over the years.

As already emphasized in Chapter 3, the distribution of studies per country is presented


in Figure 19 (left), from which stays clear the huge contribution of researchers from
Italy, Turkey and Greece to the development of analytical fragility curves. This is
naturally associated to the higher seismicity and to the high number of earthquakes that
harmed these countries in recent years, opening up the possibility of calibrating
analytical models from empirical data collected by means of post-earthquake damage
observation. From Figure 19 (centre) one can observe that 62% of the studies are
developed for existing buildings (EB) against the 38% developed for prototype buildings
(PR). Finally, even though unreinforced masonry (URM) buildings remain one of the most
common and vulnerable building typologies worldwide, approximately 62% of the
examined fragility curve studies were developed for reinforced-concrete (RC) buildings.
This might be explained due to severe damage and corresponding losses associated to
the behaviour of this particular typology in recent earthquakes, often triggered either by
lack of adequate seismic designing or building/construction detailing.

47
GR (19%)

PR URM
IT (29%) 38% 38%

RO (5%)

CH (5%)

MK (2%)
SI (2%)
EUR (2%) EB RC
62% 62%
PT (7%)

TR (26%) SP (2%)

Figure 19. Distribution of the examined studies by country (left), building type used in the analysis
(centre), and structural material (right).

Figure 20 presents the results of the evaluation carried out considering the aspects of
the capacity. From the first pie chart, in Figure 20 (a), it is possible to observe that 31%
of the studies regarding fragility curves for RC buildings do not take into account the
influence of non-structural components such as infill masonry walls. Figure 20 (b)
depicts how the examined studies treat the variability of the building height (or number
of storeys), from which one can conclude that about 38% of these studies considered
more than three different classes, 36% considered two or three different classes, and
26% considered only one class on their analyses.
In Figure 20 (c), the distribution of the results from the evaluation carried out regarding
the analysis type is presented, from which one can observe that approximately 33% of
the studies have performed NLD analysis for the computation of the seismic response of
structures. Moreover, NLS and SMM-NLS analyses were conducted in 38% and 28% of
the reviewed studies, respectively. In turn, Figure 20 (d) presents the model type
considered within each reviewed fragility curve, from which it is possible to observe that
approximately 62% have considered SDoF systems, 26% have considered reduced MDoF
systems, and only about 13% have considered MDoF systems.
Even though most of these studies do not mention it in a clear and direct way, it was
found that about 65% of the examined studies for RC buildings have not considered the
shear failure mechanism in the seismic response of structures, as illustrated in Figure 20
(e). In turn, from Figure 20 (f) one can observe that about 69% of the studies
exclusively concerning the fragility assessment of URM buildings do not include the out-
of-plane collapse mechanisms, despite the fact that recent earthquakes and research
have demonstrated that out-of-plane mechanisms are one of the most damaging failure
mechanisms for masonry structures.
Finally, Figure 20 (g) and Figure 20 (h) depict how the consideration of horizontal
diaphragms and geometric irregularities is treated in the examined studies. It is possible
to observe that in 62% and 64% of the cases studies do not take into consideration the
variability of horizontal diaphragms or geometric irregularities, respectively.

48
H (Yes)
69%

H (>3)
38%

M (2,3)
36%

L (No) L (1)
31% 26%

(a) (b)

M (Reduced MDoF) 13%


H (NLD) 33%
M (NLS) 38%
H (MDoF) 26%

L (SDoF) 62%
L (NLS-SMM/LS) 28%

(c) (d)

H (Yes)
35% H (Yes)
31%

L (No)
L (No) 69%
65%

(e) (f)

H (Yes) 38%
H (Yes)
36%

L (No)
64%
L (No) 62%

(g) (h)
Figure 20.Distribution of the examined studies regarding capacity: components for structural
assessment (a); classes of building height (b); analysis type (c); model type (d); shear failure (e);
out-of-plane mechanism (f); horizontal diaphragms (g) and geometrical irregularities (h).

49
Figure 21 (left) shows how the seismic demand is treated amongst the examined
studies. As one can perceive, the reviewed studies have determined the seismic demand
by means of real ground motion records (51%) according to the Eurocode 8
recommendations (at least 7 accelerograms) or by using code-based spectra (49%) in
equal numbers. Moreover, from Figure 21 (right), it is possible to observe that in 85% of
the studies the seismic demand is related to the specific site of the assessed structures.
Nevertheless, one should keep in mind that the remaining 15% mainly adopted real
ground motion records (>7 accelerograms).

H (>7 accelerograms) H (Related)


51% 85%

L (Code-based spectra) L (Unrelated)


49% 15%

Figure 21. Distribution of the examined studies regarding demand: representation of seismic
demand (left), and consideration of site-specific seismic demand input data (right).

With respect to the general aspects of fragility curves, Figure 22 presents the
distribution of the results in what regards the damage states thresholds, IMs, sample
size and sampling method subcategories. Figure 22 (left) depicts how the damage state
thresholds were defined, from which it is possible to observe that approximately 72% of
the examined studies have considered a custom definition for the construction of the
respective fragility curves, against 28% that used pre-defined damage state thresholds.
Moreover, from Figure 22 (centre) it is possible to observe that about 90% of the
examined studies generate the fragility curves as a function of a single IM, which is in
contradiction the recommendations of several authors (Baker and Cornell 2005; Kafali
and Grigoriu 2007; Seyedi et al. 2010; Koutsourelakis 2010).
In Figure 22 (right) is presented the distribution of the size (or scale) of the building
sample considered in each of the examined studies. It can be observed that roughly
56% of these studies considered a sample of several (hundreds) of buildings, 23%
considered a sample composed by few (dozens) of buildings, and 21% have carried out
the analysis considering only one single structure.

H (Custom)
72% H (Hundreds)
23%

L (1)
90%

M (2,3)
5%
L (One)
H (>3) M (Few) 21%
5% 56%

L (Pre-set)
28%

Figure 22. Distribution of the examined studies regarding the damage states thresholds definition
(left), IMs (centre), and sample size (right).

50
Finally, Figure 23 depicts whether uncertainties related to capacity, demand and damage
state thresholds were addressed or not in the examined studies. From Figure 23 (left) it
is possible to observe that 82% of the studies have considered uncertainties in capacity
on their analyses. These can be related to geometrical, mechanical, structural or
modelling parameters. From Figure 23 (centre), one can observe that about 15% of the
examined studies have not considered the uncertainty in the seismic demand. Lastly,
Figure 23 (right) shows that approximately 31% of studies have not considered the
uncertainty in the definition of damage state thresholds.

L (No)
L (No) 15% H (Yes)
18% 69%

H (Yes) L (No)
82% H (Yes) 31%
85%

Figure 23. Distribution of the examined studies regarding the uncertainty in capacity (left),
demand (centre), and damage state thresholds (right).

51
52
6. Concluding remarks

This final chapter presents a general overview of the work developed throughout this
report, which was focused on reviewing existing analytical fragility curves suitable for the
European building stock. To conclude, possible future research lines are proposed.
In Chapter 2, a brief overview of the existing approaches – namely analytical, empirical,
based on expert judgement and hybrid – for the derivation of fragility curves was
presented and the main advantages and drawbacks of the most common methodologies
for each type of approach were discussed.
In Chapter 3, particular attention was given to analytical methodologies for deriving
fragility curves. A comprehensive literature review was carried out focusing exclusively
on the methodologies for deriving fragility curves developed specifically for the European
building stock. Additionally, the most relevant factors that influence the reliability of the
reviewed analytical methodologies were carefully addressed and discussed, in order to
provide a clear insight about the existing literature.
Moreover, in Chapter 4, these factors were taken into account for the establishment of
evaluation criteria, aiming to provide valuable support on the selection of the most
appropriate set of fragility curves for a given geographical location and structural
typology.
Finally, in Chapter 5, the main trends of existing fragility curves for the European
building stock were identified, by carrying out a statistical analysis of the evaluation
criteria outputs of all the subcategories described in the previous chapters.
It is important to stress out that in this report only analytical methodologies for deriving
fragility curves were addressed. It would be interesting to extend this study to empirical,
expert judgment elicitation and hybrid approaches. Moreover, it would be equally
interesting to extend this work beyond the European context. The evaluation criteria
herein proposed, though aiming at a qualitative assessment, are a starting point for
further research on how fragility curves should be ideally derived, and a useful guide to
support the decisions regarding the selection of methodologies and approaches for the
generation of fragility curves.
Additionally, given the large number of methodologies that have been developed and
applied in the last decades for developing analytical fragility curves, guidelines for
the future development of these curves are needed. In this way, the selection of
curves by non-expert users will be much easier, promoting the use of seismic risk
assessment in risk mitigation policymaking. In order to improve existing fragility
curves, the following challenges should be further investigated:
 The consideration of shear failure should be taken into account when deriving
fragility curves not only in URM buildings but also in RC structures (elements);
 The uncertainty in the definition of damage states thresholds should be equally
considered when deriving fragility curves;
 Fragility curves should be developed for other building types, such as prefabricated
structures of variable size and use, as tackled by Cassotto et al. (2015);
 The impact of aging and progressive deterioration of the material properties caused
by aggressive environmental conditions, as for example the corrosion of RC
structural elements due to chloride penetration, which induce significant variation of
the mechanical properties, should be taken into consideration, as attempted by
Pitilakis et al. (2014);
 The effect of soil-structure interaction should be considered in the derivation of
fragility functions for RC buildings, as tackled for instance in Pitilakis et al. (2014);
 Existing fragility curves should be validated with empirical and experimental data.

53
54
References

1. Abbasnia R, Davoudi AT and Maddah MM (2013) An adaptive pushover procedure based on


effective modal mass combination rule. Engineering Structures, 52:654–666.
doi:10.1016/j.engstruct.2013.03.029.
2. Ahmad N, Crowley H and Pinho R (2011) Analytical Fragility Functions for Reinforced
Concrete and Masonry Buildings and Buildings Aggregates of Euro-Mediterranean Regions
– UPAV methodology. Internal Report, SYNER-G Project 2009/2012.
3. Akansel VH, Yakut A and Gulkan P (2012) Fragility of shear wall buildings with torsional
Irregularity. In Proceedings of the 15th World Conference on Earthquake Engineering
(WCEE). Lisbon, Portugal.
4. Akkar S, Sucuoğlu H and Yakut A (2005) Displacement-based fragility functions for low-
and mid-rise ordinary concrete buildings. Earthquake Spectra, 21(4):901-927.
doi:10.1193/1.2084232.
5. ANSYS (2005) Guide to ANSYS User Programmable Features, ANSYS Release 10.0,
Canonsburg (PA), USA. www.ansys.com.
6. Antoniou S, Pinho R (2004) Development and verification of a displacement-based
adaptive pushover procedure. Journal of Earthquake Engineering, 8(5):643–661.
7. Aspinall W and Cooke RM (1998) Expert Judgement and the Montserrat Volcano eruption.
In Proceedings of the 4th International Conference on Probabilistic Safety Assessment and
Management (PSAM4). Vol. 3, pp. 13-18.
8. ATC (1985) Earthquake Damage Evaluation Data for California (ATC-13). Applied
Technology Council. Redwood City (CA), USA.
9. ATC (2011) Seismic performance assessment of buildings, Volume 1 – Methodology (ATC-
58). Applied Technology Council. Redwood City (CA), USA.
10. Baker JW and Cornell CA (2005) A vector-valued ground motion intensity measure
consisting of spectral acceleration and epsilon. Earthquake Engineering and Structural
Dynamics, 34(10):1193–1217. doi:10.1002/eqe.474.
11. Barbat AH, Pujades LG and Lantada N (2006) Performance of buildings under earthquakes
in Barcelona, Spain. Computer-Aided Civil and Infrastructure Engineering, 21(8):573-593.
doi:10.1111/j.1467-8667.2006.00450.x.
12. Bernardini A, D’Ayala D, Modena C, Speranza E and Valluzzi MR (2008) Vulnerability
assessment of the historical masonry building typologies of Vittorio Veneto. Bollettino di
Geofisica Teorica e Applicata, 49(3-4):463-484.
13. Bernardini A, Gori R and Modena C (1990) An Application of Coupled Analysis Models and
Experimental Knowledge for Seismic Vulnerability Analysis of Masonry Buildings. In:
Koridze A (ed) 3:161-180 Engineering Aspects of Earthquake Phenomena. Omega
Scientific, Oxon.
14. Bertero VV and Brokken S (1983) Infills in Seismic Resistant Building. Journal of Structural
Engineering (ASCE), 109(6):1337-1361.
15. Bilgin H (2013) Fragility-based assessment of public buildings in Turkey. Engineering
Structures, 56:1283–1294. doi:10.1016/j.engstruct.2013.07.002.
16. Bommer JJ and Acevedo AB (2004) The use of real earthquake accelerograms as input to
dynamic analysis, Journal of Earthquake Engineering, 8:43–91. doi:
10.1142/S1363246904001596.
17. Borzi B, Crowley H and Pinho R (2008a) The influence of infill panels on vulnerability
curves for RC buildings. In Proceedings of the 14th World Conference on Earthquake
Engineering (WCEE). Beijing, China.
18. Borzi B, Crowley H and Pinho R (2008c) Simplified pushover-based earthquake loss
assessment (SP-BELA) method for masonry buildings. International Journal of Architectural
Heritage, 2(4):353–376. doi:10.1080/15583050701828178.

55
19. Borzi B, Pinho R and Crowley H (2007) “SP-BELA: un metodo meccanico per la definizione
della vulnerabilità basato su analisi pushover semplificate”. In Proceedings of the XII
Convegno L'Ingegneria Sismica in Italia (ANIDIS). Pisa, Italy (in Italian).
20. Borzi B, Pinho R and Crowley H (2008b) Simplified pushover-based vulnerability analysis
for large-scale assessment of RC buildings. Engineering Structures, 30(3):804–820.
doi:10.1016/j.engstruct.2007.05.021.
21. Bosiljkov V, Kržan M and D’Ayala D (2012) Vulnerability study of urban and rural heritage
masonry in Slovenia through the assessment of local and global seismic response of
buildings. In Proceedings of the 15th World Conference on Earthquake Engineering (WCEE).
Lisbon, Portugal.
22. Brzev S, Scawthorn C, Charleson AW, Allen L, Greene M, Jaiswal K and Silva V (2013) GEM
Building Taxonomy Version 2.0, GEM Technical Report 2013-02 V1.0.0, pp. 188. GEM
Foundation. Pavia, Italy. doi:10.13117/GEM.EXP-MOD.TR2013.02.
23. Calvi GM (1999) A Displacement-Based Approach For Vulnerability Evaluation Of Classes Of
Buildings. Journal of Earthquake Engineering, 3(3):411-438. doi:
10.1142/S136324699900017X.
24. Calvi GM, Pinho R, Magenes G, Boomer JJ, Restrepo-Velez LF and Crowley H (2006)
Development of seismic vulnerability assessment methodologies over the past 30 years.
ISET Journal of Earthquake Technology, 43(3):75-104.
25. Casarotti C, Monteiro R and Pinho R (2009) Verification of spectral reduction factors for
seismic assessment of bridges. Bulletin of the New Zealand Society for Earthquake
Engineering, 42(2):111–121.
26. Casotto C, Silva V, Crowley H, Nascimbene R and Pinho R (2015) Seismic fragility of Italian
RC precast industrial structures. Engineering Structures, 94:122-136.
doi:10.1016/j.engstruct.2015.02.034.
27. Cattari S and Lagomarsino S (2012) Performance-based approach to earthquake protection
of masonry cultural heritage. In: Proceedings of structural analysis of historical
constructions (2012):pp. 2914–2922.
28. CEN (2004) Eurocode 8: design of structures for earthquake resistance - Part I: general
rules, seismic actions and rules for buildings. European Committee for Standardization
(CEN). Brussels, Belgium.
29. Ceran HB (2010) Seismic vulnerability of masonry structures in Turkey. Master Thesis.
Department of Civil Engineering of the Middle East Technical University, Turkey.
30. Chaulagain H, Rodrigues H, Jara J, Spacone E and Varum H (2013) Seismic response of
current RC buildings in Nepal: a comparative analysis of different design/construction.
Engineering Structures, 49:284-294.
31. Chopra AK, Goel RK (2000) Evaluation of NSP to estimate seismic deformation: SDF
systems. Journal of Structural Engineering, 126(4):482–490. 

32. Clemen RT and Winkler RL (1999) Combining Probability Distributions From Experts in Risk
Analysis. Risk Analysis, 19:187-203. doi:10.1111/j.1539-6924.1999.tb00399.x.
33. Computers and Structures Inc. (2002) SAP2000 - Integrated Software for Structural
Analysis & Design, Berkeley (CA), USA.
34. Cooke RM (1991) Experts in Uncertainty - Opinion and Subjective Probability in Science.
Environmental Ethics and Science Policy Series. Oxford University Press, New York 10016.
isbn:0-19-506465-8.
35. Cooke RM and Goossens LHJ (2000) Expert Judgment Elicitation in Risk Assessment. Delft
University of Technology, Abstract NATO workshop, Lisbon, 1-4 October 2000.
36. Cooke RM (2001) EXCALIBUR: Windows version of EXCALIBUR, software for performance
based combination of expert judgments. Department of Mathematics, Delft University of
Technology, Netherlands.
37. Cosenza E, Manfredi G, Polese M and Verderame GM (2005) A Multi-Level Approach to the
Capacity Assessment of Existing RC Buildings. Journal of Earthquake Engineering, 9:1-22.

56
38. Costa AA (2012) Seismic assessment of the out-of-plane performance of traditional stone
masonry walls. PhD Thesis. Faculty of Engineering of the University of Porto.
39. Crowley H, Pinho R and Bommer JJ (2004) A probabilistic displacement-based vulnerability
assessment procedure for earthquake loss estimation. Bulletin of Earthquake Engineering,
2:173-219.
40. D’Ayala D (2004) Correlation of Fragility Curves for Vernacular Building Types: Houses in
Lalitpur, Nepal and in Istanbul, Turkey. In Proceedings of the 13th World Conference on
Earthquake Engineering (WCEE). Vancouver (BC), Canada. August 1-6. Available at:
http://www.iitk.ac.in/nicee/wcee/article/13_485.pdf.
41. D’Ayala D (2005) Force and displacement based vulnerability assessment for traditional
buildings. Bulletin of Earthquake Engineering, 3(3):235-265. doi:10.1007/s10518-005-
1239-x.
42. D'Ayala D (2013) Assessing the seismic vulnerability of masonry buildings. In:
Tesfamariam S and Goda K (eds) Handbook of Seismic Risk management of civil
infrastructure systems. Woodhead publishing.
43. D'Ayala D, Jaiswal KS, Wald DJ, Porter K and Greene M (2010) Collaborative effort to
estimate collapse fragility for buildings worldwide: The WHE-PAGER project. Available at:
http://discovery.ucl.ac.uk/id/eprint/1369713.
44. D'Ayala D and Kansal A (2004) Analysis of the seismic Vulnerability of the architectural
Heritage in Buhj, Gujarat, India. In Modena C, Lourenço PB and Roca P (eds) Structural
Analysis of Historical Construction, Vol. 2, Possibilities of numerical and experimental
techniques, pp. 1069-1078. isbn:0415363799.
45. D’Ayala D and Kishali E (2012) Analytically derived fragility curves for unreinforced
masonry buildings in urban contexts. In Proceedings of the 15th World Conference on
Earthquake Engineering (WCEE). Lisbon, Portugal.
46. D’Ayala D and Meslem A (2012) Guide for selection of existing analytical fragility curves
and compilation of the database. GEM Technical Report 2012-X. GEM Foundation. Pavia,
Italy.
47. D’Ayala D and Meslem A (2013) Sensitivity of analytical fragility functions to capacity-
related parameters. GEM Technical Report 2013-X, GEM Foundation. Pavia, Italy.
48. D’Ayala D, Meslem A, Vamvatsikos D, Porter K, Rossetto T, Crowley H and Silva V (2014)
Guidelines for analytical vulnerability assessment of low/mid-rise buildings – methodology.
Vulnerability Global Component Project. GEM Technical Report. GEM Foundation. Pavia,
Italy.
49. D’Ayala D and Paganoni S (2011) Assessment and analysis of damage in L’Aquila historic
city centre after 6th April 2009. Bulletin of Earthquake Engineering, 9:81-104.
doi:10.1007/s10518-010-9224-4.
50. D’Ayala D, Spence R, Oliveira C and Pomonis A (1997) Earthquake loss estimation for
Europe’s historic town centres. Earthquake Spectra, 13(4):773-793.
doi:10.1193/1.1585980.
51. D’Ayala D and Speranza E (2003) Definition of collapse mechanisms and seismic
vulnerability of historic masonry buildings. Earthquake Spectra, 19(3):479-509.
doi:10.1193/1.1599896.
52. D'Ayala D and Yeomans D (2004) Assessing The Seismic Vulnerability of Late Ottoman
Buildings in Istanbul. In Proceedings of the IV International Seminar on Structural Analysis
of Historical Constructions - SAHC04. Padua, Italy.
53. Dalkey N (1969) The Delphi Method: An experimental study of group opinion. A report
prepared for United States Air Force Project Rand RM-5888-PR.
54. De Luca F, Verderame GM, Gómez-Martínez F and Pérez-García A (2014) The structural
role played by masonry infills on RC building performances after the 2011 Lorca, Spain,
earthquake. Bulletin of Earthquake Engineering, 12(5):1999-2026. doi:10.1007/s10518-
013-9500-1.

57
55. Del Gaudio C, Ricci P, Verderame GM and Manfredi G (2015) Development and urban-scale
application of a simplified method for seismic fragility assessment of RC buildings.
Engineering Structures, 91:40-57. doi:10.1016/j.engstruct.2015.01.031.
56. Delbecq A, Van de Ven A and Gusstafson D (1975) The Delphi Technique. In: Group
Techniques for Program Planning. Glenview (Illinois): Scott, Foresman and Company.
Chapter 2, pp. 83-107.
57. Dolšek M ans Fajfar P (2008) The effect of masonry infills on the seismic response of a
four-storey reinforced concrete frame - a deterministic assessment. Engineering
Structures, 30(7):1991-2001. doi:10.1016/j.engstruct.2008.01.001.
58. Dolšek M and Fajfar P (2004) Inelastic spectra for infilled reinforced concrete frames.
Earthquake Engineering and Structural Dynamics, 33(15):1395-1416.
doi:10.1002/eqe.410.
59. Dumova-Jovanoska E (2000) Fragility curves for reinforced concrete structures in Skopje
(Macedonia) region. Soil Dynamics and Earthquake Engineering, 19(6):455-466.
doi:10.1016/S0267-7261(00)00017-8.
60. Eliashberg J and Winkler RL (1981) Risk Sharing and Group Decision Making. Management
Science, 27(11):1221-1235. doi:10.1287/mnsc.27.11.1221.
61. Erberik, MA (2008) Generation of fragility curves for Turkish masonry buildings considering
in-plane failure modes. Earthquake Engineering and Structural Dynamics, 37:387-405. doi:
10.1002/eqe.760.
62. Erberik MA, Elnashai SA (2004) Fragility analysis of flat-slab structures. Engineering
Structures, 26:937-948. doi:10.1016/j.engstruct.2004.02.012.
63. Faber MH and Stewart MG (2003) Risk assessment for civil engineering facilities: critical
overview and discussion. Reliability Engineering & System Safety, 80(2):173-184, ISSN
0951-8320. doi:10.1016/S0951-8320(03)00027-9.
64. Fajfar P and Gašperšič P (1996) The N2 method for the seismic damage analysis of RC
buildings. Earthquake Engineering and Structural Dynamics, 25:31-46.
doi:10.1002/(SICI)1096-9845(199601)25:1<31::AID-EQE534>3.0.CO;2-V.
65. Fajfar P (1999) Capacity spectrum method based on inelastic demand spectra. Earthquake
Engineering and Structural Dynamics, 28:979-993.
66. Fajfar P (2000) A Nonlinear Analysis Method for Performance ‐ Based Seismic Design.
Earthquake Spectra, 16(3):573-592. doi:10.1193/1.1586128.
67. Fardis MN, Papailia A and Tsionis G (2012) Seismic fragility of RC framed and wall-frame
buildings designed to the EN-Eurocodes. Bulletin of Earthquake Engineering, 10(6):1767-
1793. doi:10.1007/s10518-012-9379-2.
68. FEMA (2003a) Multi-hazard loss estimation methodology: earthquake model (HAZUS-MH-
MR3). Technical report. Department of Homeland Security. Washington (DC), USA.
Prepared by the Federal Emergency Management Agency under a contract with the
National Institute of Building Sciences.
69. FEMA (2003b) Preliminary Evaluation of Methods for Defining Performance. Applied
Technology Council (ATC-58-2). Washington (DC), USA. Available at:
https://www.civil.iitb.ac.in/~p0saurabhrs/ATC-58-2.pdf.
70. FEMA (2008) Quantification of Building Seismic Performance Factors. Federal Emergency
Management Agency (FEMA P-695), Applied Technology Council (ATC-63). Redwood City
(CA), USA.
71. FEMA (2012) Seismic Performance Assessment of Buildings, Federal Emergency
Management Agency (FEMA P-58), Applied Technology Council (ATC-58). Washington
(DC), USA.
72. Ferreira TM (2015) Out-of-plane seismic performance of stone masonry walls:
experimental and analytical assessment. PhD Thesis. Department of Civil Engineering of
the University of Aveiro. Portugal.
73. Ferreira TM, Maio R, Vicente R (2015) Seismic vulnerability assessment of traditional
masonry buildings: the relevance of effective seismic strengthening policies. In

58
Proceedings of the International Conference on Recent Advances in Rehabilitation and
Sustainability of Structures (RehabStructures 2015), Session V: Non-structural
rehabilitation of constructions, pp. 347-361. 1-2 June 2015, University of Azores, Ponta
Delgada (Azores), Portugal.
74. Fiorato AE, Sozen MA, Gamble WL (1970) Investigation of the interaction of reinforced
concrete frames with masonry filler walls. Report UILU, Structural Research Series 370,
University of Illinois, USA.
75. Florio G (2010) Vulnerability of historical masonry buildings under exceptional actions. PhD
Thesis. Department of Structural Engineering of the University of Naples Federico II.
Available at: http://www.fedoa.unina.it/8292/.
76. Franchin P (2013) Methodology for systemic seismic vulnerability assessment of buildings,
infrastructures, networks and socio-economic impacts. SYNER-G Reference Report 1. JRC
Scientific and Policy Reports. doi:10.2788/69238.
77. Freeman SA (1998) Development and use of capacity spectrum method. In Proceedings of
the 6th US NCEE Conference on Earthquake Engineering/EERI. Seattle, Washington, USA.
78. Gambarotta L and Lagomarsino S (1996) On Dynamic Response of Masonry Panels, In
Proceedings of the National Conference Masonry Mechanics Between Theory and Practice.
Messina, Italy.
79. Gehl P, Seyedi DM and Douglas J (2012) Vector-Valued Fragility Functions for Seismic Risk
Evaluation. Bulletin of Earthquake Engineering, 11(2):365–384. doi:10.1007/s10518-012-
9402-7.
80. Genest C and Zidek JV (1986) Combining Probability Distributions: A Critique and an
Annotated Bibliography. Statistical Science, 1:114-135. Available at:
http://www.jstor.org/stable/2245510.
81. Glasiter S and Pinho R (2003) Development of A Simplified Deformation-Based Method for
Seismic Vulnerability Assessment. Journal of Earthquake Engineering, 7:107-140.
doi:10.1142/S1363246903000985.
82. Goicoechea A, Hansen DR and Duckstein L (1982) Multiobjective Decision Analysis with
Engineering and Business Applications. New York: John Wiley and Sons. isbn:13:978-
0471064015.
83. Grünthal G (ed) (1998) European Macroseismic Scale 1998 (EMS-98) European
Seismological Commission, sub commission on Engineering Seismology, Working Group
Macroseismic Scales. Conseil de l’Europe, Cahiers du Centre Européen de Géodynamique
et de Séismologie, Vol. 15, Luxembourg.
84. Gündüz A (1996) Probability, statistics, risk and reliability in engineering. Kure
publications, Istanbul, Turkey. 

85. Gustafson D, Shulka R, Delbecq A and Walster A (1973) A Comparative Study of
Differences in Subjective Likelihood Estimates Made by Individuals, Interacting Groups,
Delphi Groups, and Nominal Groups. Organizational Behaviour and Human Performance,
9:280-291. doi:10.1016/0030-5073(73)90052-4.
86. Hancilar U and Taucer F (2013) Guidelines for typology definition of European physical
assets for earthquake risk assessment. SYNER-G Reference Report 2. JRC Scientific and
Policy Reports. doi:10.2788/68751.
87. Hancilar U and Çaktr E (2015) Fragility functions for code complying RC frames via best
correlated IM–EDP pairs. Bulletin of Earthquake Engineering, 13(11):3381-3400.
doi:10.1007/s10518-015-9775-5.
88. IDARC-2D (1996) IDARC-2D V5.0 Computer program - Inelastic damage analysis of RC
building structures. Developed by Park YJ, Reinhorn AM and Kunnath SK. State University
of New York, USA.
89. Iervolino I, Maddaloni G and Cosenza E (2008) Eurocode 8 Compliant Real Record Sets for
Seismic Analysis of Structures. Journal of Earthquake Engineering, 12:54-90. doi:
10.1080/13632460701457173.
90. Indirli M, Clemente P, Spadoni B, Cami R, Speranza E, Mucciarella M and Pistola F (2004)
Seismic protection of historical centers using innovative techniques, with focus on San

59
Giuliano di Puglia after the 2002 Molise earthquake. In Proceedings of the Structural
analysis of historical construction (SAHC), IV International Seminar. Padua, Italy.
91. Jaiswal K and Wald D (2010) An Empirical Model for Global Earthquake Fatality Estimation.
Earthquake Spectra, 26(4):1017-1037. doi:10.1193/1.3480331.
92. Jaiswal K, Wald D and D’Ayala D (2011) Developing Empirical Collapse Fragility Functions
for Global Building Types. Earthquake Spectra, 27(3):775-795. doi:
http://dx.doi.org/10.1193/1.3606398.
93. Jaiswal KS, Aspinal WP, Perkins D, Wald D and Porter KA (2012) Use of Expert Judgment
Elicitation to Estimate Seismic Vulnerability of Selected Building Types. In Proceedings of
the 15th World Conference on Earthquake Engineering (WCEE). Lisbon, Portugal.
94. Jiang H, Lu X and Chen L (2012) Seismic Fragility Assessment of RC Moment-Resisting
Frames Designed According to the Current Chinese Seismic Design Code. Journal of Asian
Architecture and Building Engineering, 11:153-160. doi:10.3130/jaabe.11.153.
95. Kafali C and Grigoriu M (2007) Seismic fragility analysis: application to simple linear and
nonlinear systems. Earthquake Engineering and Structural Dynamics, 36(13):1885–1900.
doi:10.1002/eqe.726.
96. Kaynia AM (2013) Guidelines for deriving seismic fragility functions of elements at risk:
Buildings, lifelines, transportation networks and critical facilities. SYNER-G Reference
Report 4. JRC Scientific and Policy Reports. doi:10.2788/19605.
97. Kappos A and Dymiotis C (2000) DRAIN-2000: a program for the inelastic time-history and
seismic reliability analysis of 2-D structures. Edinburgh, UK: Department of Civil and
Offshore Engineering, Heriot Watt University, Report no. STR/00/CD/01.
98. Kappos A, Lekidis V, Panagopoulos G, Sous I, Theodulidis N, Karakostas C, Margaris B
(2007) Analytical estimation of economic loss for buildings in the area struck by the 1999
Athens earthquake and comparison with statistical repair costs. Earthquake Spectra,
23(2):333-355. doi:10.1193/1.2720366.
99. Kappos A, Panagiotopoulos C, Panagopoulos G and Papadopoulos E (2003). RISK-UE: an
advanced approach to earthquake risk scenarios with applications to deffierent European
towns. WP4 – reinforced concrete buildings (level I and II analysis). Thessaloniki, Greece.
100. Kappos A and Panagopoulos G (2009) Fragility curves for reinforced concrete buildings in
Greece. Structure and Infrastructure Engineering, 6(1-2):39-53.
doi:10.1080/15732470802663771.
101. Kappos A, Panagopoulos G, Panagiotopoulos C and Penelis G (2006) A hybrid method for
the vulnerability assessment of RC and URM buildings. Bulletin of Earthquake Engineering,
4(4):391-413. doi:10.1007/s10518-006-9023-0.
102. Kappos A, Stylianidis KC and Pitilakis K (1998) Development of seismic risk scenarios
based on a hybrid method of vulnerability assessment. Natural Hazards, 17(2):177-192.
doi:10.1023/A:1008083021022.
103. Karantoni F, Tsionis G, Lyrantzaki F and Fardis MN (2014) Seismic fragility of regular
masonry buildings for in-plane and out-of-plane failure. Earthquake and Structures,
6(6):689-713. doi:10.12989/eas.2014.
104. Kircil MS and Polat Z (2006) Fragility analysis of RC frame buildings on firm sites. In
Proceedings of the 2nd International Congress Fédération Internationale du Béton. Naples,
Italy.
105. Klingner RE and Bertero VV (1976) Infilled frames in earthquake-resistant construction.
Report 76-32, University of California, Berkeley (CA), USA.
106. Klügel J (2011) Uncertainty Analysis and Expert Judgment in Seismic Hazard Analysis. Pure
and Applied Geophysics, 168(1-2):27-53. doi:10.1007/s00024-010-0155-4.
107. Koutsourelakis P (2010) Assessing structural vulnerability against earthquakes using multi-
dimensional fragility surfaces: a bayesian framework. Probabilistic Engineering Mechanics,
25:49–60. doi:10.1016/j.probengmech.2009.05.005.

60
108. Lagomarsino S and Cattari S (2014) Fragility functions of masonry buildings. In SYNER-G:
Typology definition and fragility functions for physical elements at seismic risk. Springer
Science & Business Media Dordrecht. doi:10.1007/978-94-007-7872-6_5.
109. Lagomarsino S and Giovinazzi S (2006) Macroseismic and Mechanical Models for the
Vulnerability and Damage Assessment of Current Buildings. Bulletin of Earthquake
Engineering, 4(4):415-443. doi:10.1007/s10518-006-9024-z.
110. Lang K (2002) Seismic vulnerability of existing buildings. PhD Thesis. Institute of
Structural Engineering of the Swiss Federal Institute of Technology. Switzerland.
111. Liel AB, Haselton CB, Deierlein GG and Baker JW (2009) Incorporating modeling
uncertainties in the assessment of seismic collapse risk of buildings. Structural Safety,
31(2):197-211. doi:10.1016/j.strusafe.2008.06.002.
112. Lin YY, Chang KC and Wang YL (2004) Comparison of displacement coefficient method and
capacity spectrum method with experimental results of RC columns. Earthquake
Engineering and Structural Dynamics, 33:35–48.
113. Lourdes AM, López RR and Saffar A (2007) Development of fragility curves for medium rise
reinforced concrete shear wall residential buildings in Puerto Rico. In Elaskar SA, Pilotta EA
and Torres GA (eds) Mecânica Computacional Vol XXVI, pp.2712-2727. Córdoba,
Argentina.
114. Lourenço PB, Mendes N, Ramos LF and Oliveira DV (2011) Analysis of Masonry Structures
Without Box Behavior. International Journal of Architectural Heritage, 5(4-5). doi:
10.1080/15583058.2010.528824.
115. Luco N and Cornell CA (2007) Structure-Specific Scalar Intensity Measures for Near-Source
and Ordinary Earthquake Ground Motions. Earthquake Spectra, 23(2):357-392. doi:
http://dx.doi.org/10.1193/1.2723158.
116. Luechinger P and Fischer J (2015) New European Techincal Rules for the Assessment and
Retrofitting of Existing Structures. JRC Science and Policy Reports. doi:10.2788/052881.
117. Maio R, Vicente R, Formisano A and Varum H (2015) Seismic vulnerability of building
aggregates through hybrid and indirect assessment techniques. Bulletin of Earthquake
Engineering. doi:10.1007/s10518-015-9747-9.
118. Mehrabi AB, Shing PB, Schuller MP and Noland JL (1994) Performance of Masonry-Infilled
RC Frames Under In-Plane Lateral Loads. Report No. CU/SR-94-6, Department of Civil,
Environmental and Architectural Engineering, University of Colorado, Boulder (CO), USA.
119. Morris PA (1997) Combining expert judgments. A Bayesian approach. Management
Science, 23(7):679-693. doi:10.1287/mnsc.23.7.679.
120. Mosleh A, Bier VM and Apostolakis G (1988) A Critique of Current Practice for the Use of
Expert Opinions in Probabilistic Risk Assessment. Reliability Engineering and System
Safety, 20:63-85. doi:10.1016/0951-8320(88)90006-3.
121. Munari M and Valluzzi MR (2009) Classificazioni di vulnerabilità dal calcolo limite per
macroelementi: applicazione ad aggregati edilizi in muratura in centri storici umbri. In
Proceedings of the 13th Conference of the Italian National Association of Earthquake
Engineering (ANIDIS). Bologna, Italy.
122. NIBS-FEMA (2003) Multi-hazard loss estimation methodology - earthquake model.
Advanced engineering building module - Technical and user’s manual. National Institute of
Building Sciences, Federal Emergency Management Agency. Washington (DC), USA.
123. Novelli VI, D’Ayala D, Makhloufi N, Benouar D and Zekagh A (2015) A procedure for the
identification of the seismic vulnerability at territorial scale. Application to the Casbah of
Algiers. Bulletin of Earthquake Engineering, 13:177–202. doi:10.1007/s10518-014-9666-
1.
124. NTC (2008) Decreto Ministeriale 14/1/2008 - Norme Tecniche per Le Costruzioni. Ministry
of Infrastructures and Transportations, Gazzetta Ufficiale S.O. no.30 on 4/2/2008 (2008).
Italy. doi:10.1515/9783110247190.153.
125. OpenSEES (2015) http://opensees.berkeley.edu/.

61
126. Oropeza M, Michel C, Bigler M and Lestuzzi P (2010) New analytical fragility curves for
existing URM buildings in regions with moderate seismicity. In Proceedings of the 8th
International Masonry Conference. Dresden, Germany.
127. Ouchi F (2004) A Literature Review on the Use of Expert Opinion in Probabilistic Risk
Analysis. World Bank Policy Research Working Paper 3201.
128. Özer AY and Erberik MA (2008) Vulnerability of Turkish low-rise and mid-rise reinforced
concrete frame structures. Journal of Earthquake Engineering, 12(2):2–11.
doi:10.1080/13632460802012687.
129. Pagnini LC, Vicente R, Lagomarsino S and Varum H (2011) A mechanical model for the
seismic vulnerability assessment of old masonry buildings. Earthquakes and Structures,
2:25-42. Available at: http://technopress.kaist.ac.kr/?journal=eas&subpage=5.
130. Papanikolaou VK, Elnashai AS and Pareja JP (2005) Limits of applicability of conventional
and adaptive pushover analysis for seismic response assessment. Mid-America Earthquake
Center, Civil and Environmental Engineering Department, University of Illinois at Urbana-
Champaign, USA.
131. Parenté FJ and Anderson-Parenté JK (1987) Delphi inquiry systems. In: Wright G and
Ayton P (eds) Judgmental Forecasting. Wiley. Chichester.
132. Park Y and Ang A (1985). Mechanistic Seismic Damage Model for Reinforced Concrete.
Journal of Structural Engineering, 111(4):722-739. doi:10.1061/(ASCE)0733-
9445(1985)111:4(722).
133. Pasticier L, Amadio C and Fragiacomo M (2008) Non-linear seismic analysis and
vulnerability evaluation of a masonry building by means of the SAP2000 V.10 code.
Earthquake Engineering and Structural Dynamics, 37:467-485. doi:10.1002/eqe.
134. Paulay T and Priestly MJN (1992) Seismic design of reinforced concrete and masonry
buildings. Hoboken (NJ), USA. John Wiley & Sons, Inc.
135. Penna A (2002) A Macro-Element Procedure for the Non-Linear Dynamic Analysis of
Masonry Buildings. PhD Thesis. Politecnico di Milano. Milan, Italy. (in Italian).
136. Pinho R, Bommer JJ and Glaister S (2002) A simplified approach to displacement-based
earthquake loss estimation analysis. In Proceedings of the 12th European Conference on
Earthquake Engineering (ECEE). England.
137. Pitilakis KD, Argyroudis S, Kakderi K and Argyroudi A (2013) Systemic seismic vulnerability
and risk analysis for buildings, lifeline networks and infrastructures safety gain. SYNER-G
synthetic document. JRC Scientific and Policy Reports. doi:10.2788/23242.
138. Pitilakis KD, Crowley H and Kaynia AM (ed.) (2014a) SYNER-G : Typology Definition and
Fragility Functions for Physical Elements at Seismic Risk. Springer Dordrecht Heidelberg
New York London. doi:10.1007/978-94-007-7872-6.
139. Pitilakis KD, Karapetrou ST and Fotopoulou SD (2014b) Consideration of aging and SSI
effects on seismic vulnerability assessment of RC buildings. Bulletin of Earthquake
Engineering, 12(4): 1755-1776. doi:10.1007/s10518-013-9575-8.
140. Plous S (1993) The Psychology of Judgment and Decision Making. New York, McGraw-Hill.
isbn-13: 978-0070504776.
141. Polese M, Verderame GM, Mariniello C, Iervolino I and Manfredi G (2008) Vulnerability
analysis for gravity load designed RC buildings in Naples - Italy. Journal of Earthquake
Engineering, 12(S2):234-245. doi:10.1080/13632460802014147.
142. Porter KA, Farokhnia K, Cho I, Grant DN, Jaiswal K, Wald D and Noh H (2012) Global
vulnerability estimation methods for the Global Earthquake Model. In Proceedings of the
15th World Conference on Earthquake Engineering (WCEE). Lisbon, Portugal.
143. Porter KA, Hu Z, Huyck C and Bevington J (2014) User guide: field sampling strategies for
estimating building inventories. GEM Foundation.
144. Restrepo-Velez LF and Magenes G (2004) Simplified procedure for the seismic risk
assessment of unreinforced masonry buildings. In Proceedings of the 13th World
Conference on Earthquake Engineering (WCEE). Vancouver, Canada.

62
145. Rossetto T, D’Ayala D, Ioannou I and Meslem A (2014). Evaluation of existing fragility
curves. In SYNER-G: Typology definition and fragility functions for physical elements at
seismic risk. Springer Science & Business Media Dordrecht. doi:10.1007/978-94-007-
7872-6_3.
146. Rossetto T and Elnashai A (2005) A new analytical procedure for the derivation of
displacement-based vulnerability curves for populations of RC structures. Engineering
Structures, 27(3):397-409. doi:10.1016/j.engstruct.2004.11.002.
147. Rossetto T, Ioannou I and Grant DN (2013). Existing empirical fragility and vulnerability
relationships: compendium and guide for selection. GEM Technical Report 2013-X. GEM
Foundation. Pavia, Italy.
148. Rota M, Penna A and Magenes G (2010) A methodology for deriving analytical fragility
curves for masonry buildings based on stochastic nonlinear analyses. Engineering
Structures, 32(5):1312-1323. doi:10.1016/j.engstruct.2010.01.009.
149. Rota M, Penna A, Strobbia C and Magenes G (2011) Typological seismic risk maps for Italy.
Earthquake Spectra, 27(3):907-926.
150. Rota M, Penna A and Magenes G (2014) A framework for the seismic assessment of
existing masonry buildings accounting for different sources of uncertainty. Earthquake
Engineering and Structural Dynamics, 43:1045-1066. doi:10.1002/eqe.2386
151. Sackman H (1975) Delphi Critique: Expert Opinion, Forecasting and Group Processes.
Lexington Books. Lexington (MA), USA.
152. Scheibe M, Skutsch M and Schofer J (1975) Experiments in Delphi methodology. In:
Linestone H and Turoff M (eds) The Delphi method: techniques and application. Addison-
Wesley. London, England.
153. Senaldi I, Magenes G, Penna A, Galasco A and Rota M (2014) The Effect of Stiffened Floor
and Roof Diaphragms on the Experimental Seismic Response of a Full-Scale Unreinforced
Stone Masonry Building. Journal of Earthquake Engineering, 18(3): 407-443.
doi:10.1080/13632469.2013.876946.
154. Seyedi DM, Gehl P, Douglas J, Davenne L, Mezher N and Ghavamian S (2010)
Development of seismic fragility surfaces for reinforced concrete buildings by means of
nonlinear time-history analysis. Earthquake Engineering and Structural Dynamics, 39:91–
108. doi:10.1002/eqe.939.
155. Silva V, Crowley H, Varum H and Pinho R (2014) Seismic risk assessment for mainland
Portugal. Bulletin of Earthquake Engineering. doi:10.1007/s10518-014-9630-0.
156. Silva V, Crowley H, Varum H, Pinho R and Sousa R (2013) Evaluation of analytical
methodologies used to derive vulnerability functions. Earthquake Engineering and
Structural Dynamics, 43(2):1-24. doi:10.1002/eqe.2337.
157. Simões A, Milošević J, Meireles H, Bento R, Cattari S and Lagomarsino S (2015) Fragility
curves for old masonry building types in Lisbon. Bulletin of Earthquake Engineering.
doi:10.1007/s10518-015-9750-1.
158. Sniezek J (1992) Groups under uncertainty: An Examination of Confidence in Group
Decision Making. Organizational Behavior and Human Decision Processes, 52:124-155.
doi:10.1016/0749-5978(92)90048-C.
159. STADATA (2011) 3Muri: seismic analysis program for 3D masonry buildings, Release 5.0.4.
Turin, Italy.
160. Tomaževič M, Weiss P and Velechovsky T (1991) The influence of rigidity of floors on the
seismic behaviour of old stone-masonry buildings. European Earthquake Engineering,
3:28-41.
161. Tsimplokoukou K, Lamperti M and Negro P (2014) Building Design for Safety and
Sustainability. JRC Science and Policy Reports. doi:10.2788/338223.
162. Tsionis G, Apostolska R and Taucer F (2014) Seismic strengthening of RC buildings. JRC
Science and Policy Reports. doi:10.2788/138156.

63
163. Tsionis G and Fardis MN (2014) Seismic fragility curves for reinforced concrete buildings
and bridges in Thessaloniki. In Proceedings of the 2nd European Conference on Earthquake
Engineering and Seismology (ECEES). Istanbul, Turkey.
164. Vacareanu R, Chesca AB, Georgescu B and Seki M (2007) Case study on the expected
seismic losses of soft and weak groundfloor buildings. In Proceedings of the International
Symposium on Strong Vrancea Earthquakes and Risk Mitigation. Bucharest, Romania.
165. Vacareanu R, Radoi R, Negulescu C and Aldea A (2004) Seismic vulnerability of RC
buildings in Bucharest, Romania. In Proceedings of the 13th World Conference on
Earthquake Engineering (WCEE). Vancouver, Canada.
166. Verderame GM, De Luca F, Ricci P and Manfredi G (2011) Preliminary analysis of a soft-
storey mechanism after the 2009 L'Aquila earthquake. Earthquake Engineering and
Structural Dynamics, 40(8):925–944. doi:10.1002/eqe.1069.
167. Von Winterfeld D (1989) Eliciting and Communicating Expert Judgments: Methodology and
Application to Nuclear Safety. Joint Research Centre, Commission of the European
Communities.
168. Wen YK, Ellingwood BR and Bracci J (2004) Vulnerability Function Framework for
Consequence-Based Engineering, Mid-America Earthquake Center, CD Release 04-04.

169. Winker RL, Hora SC and Baca RG (1992) The Quality of Experts’ Probabilities Obtained
Through Formal Elicitation Techniques. Center for Nuclear Waste Regulatory Analyses
CNWRA. San Antonio (TX), USA.

170. Zahedi F (1986) Group Consensus Function Estimation when Preferences are Uncertain.
Operations Research, 34(6):883-894. doi:10.1287/opre.34.6.883.

64
List of abbreviations and symbols

ATC, Applied Technology Council


AI, Arias Intensity
CL, Continuous Lognormal Distribution
CN, Continuous Normal Distribution
CSM, Capacity Spectrum Method
DCM, Displacement Coefficient Method
DD, Discrete Distribution
DLS, Damage Limitation States
DPM, Damage Probability Matrix
EB, Existing Building
EDP, Engineering Demand Parameter
EEFIT, Earthquake Engineering Field Investigation Team
EERI, Earthquake Engineering Research Institute
ELSA, European Laboratory for Structural Assessment
EU, European Union
GD, Global Drift
GEM, Global Earthquake Model
GMPE, Ground Motion Prediction Equation
GVC, Global Vulnerability Consortium
IDA, Incremental Dynamic Analysis
IM, Intensity Measure
IN2, Incremental N2 Method
ISD, Inter-storey Drift
JRC, Joint Research Centre
LHS, Latin Hypercube Sampling
LS, Linear Static
MDoF, Multi Degree of Freedom
MFA, Maximum Floor Acceleration
MMI, Modified-Mercalli Intensity
MPR, Maximum Plastic end Rotation of beams
Mw, Moment magnitude
NC, Near Collapse
NLD, Non-linear Dynamic Analysis
NLS, Non-linear Static Analysis
PR, Prototype (building)
RC, Reinforce Concrete
RD, Roof Displacement
rrup, rupture distance
SAPA, Static Adaptive Pushover Analysis
Sa(T), Spectral acceleration
SD, Significant Damage
Sd(T), Spectral displacement
SDoF, Single Degree of Freedom
SMM, Simplified Mechanical Methods
Sv(T), Spectral velocity
TD, Top Displacement
UD, Ultimate Displacement
URM, Unreinforced Masonry
ULS, Ultimate Limit States
VD, Median base shear demand given the PGA
VLS,i, Median base shear capacity at the i th limit state
vs,30, Average shear-wave velocity between 0 and 30 m

65
66
List of figures

Figure 1. Flow diagram for deriving analytical fragility curves. .................................... 6


Figure 2. Distribution of the examined analytical fragility assessment methodologies by
country: Italy (12); Turkey (11); Greece (8); Portugal (3); Romania (2); Switzerland
(2); former Yugoslav Republic of Macedonia (1); Slovenia (1) and Spain (1). ............. 12
Figure 3. Main components and phases considered in analytical fragility assessment
methodologies and associated uncertainties. .......................................................... 13
Figure 4. Example of a masonry structural wall system (left) and the corresponding
three-dimensional assemblage of all the structural walls (right), modelled through a
macro-element approach using the 3Muri ® software (Ferreira et al. 2015). ................ 14
Figure 5. Examples of reinforced concrete moment-resisting frame (left) and dual frame-
wall structural systems (right) (Dumova-Jovanoska 2000). ...................................... 15
Figure 6. Numerical representation of infill masonry walls through equivalent diagonal
strut model (Amato et al. 2008). ........................................................................... 16
Figure 7. Comparison between fragility curves for 4-storey RC buildings with and without
considering masonry infill walls in structural modelling (D’Ayala and Meslem 2012). .... 17
Figure 8. IDA curve with the identification of yielding and collapse points (Kirçil and Polat
2006). ............................................................................................................... 18
Figure 9. Example of (bilinear) capacity curves derived from conventional pushover
analysis (left) and those obtained from using a displacement-based adaptive pushover
(right), developed by Ferreira et al. (2015) and Silva et al. (2013), respectively. ........ 20
Figure 10. Out-of-plane failure mechanisms of the FaMIVE method developed by D’Ayala
and Speranza (2003). .......................................................................................... 22
Figure 11. Flowchart of the DBELA method (Silva et al. 2013). ................................. 23
Figure 12. Flowchart of the SP-BELA method (Borzi et al. 2008b). ............................. 24
Figure 13. Vulnerability curves for 4-storey buildings using the SP-BELA and DBELA
methodologies for Limit State 2 (left) and Limit State 3 (right) conditions (Borzi et al.
2008b). ............................................................................................................. 25
Figure 14. Comparison of capacity curves obtained by the three-dimensional model with
those obtained by superposition of two-dimensional models (D’Ayala and Meslem 2013).
......................................................................................................................... 26
Figure 15. Configuration and dimensions of typical four-storey irregularly infilled RC
frame building (Kappos and Panagopoulos 2009). ................................................... 27
Figure 16. Existing techniques for assessing the out-of-plane behaviour of unreinforced
masonry structures (Ferreira 2015). ...................................................................... 29
Figure 17. Continuous fragility curves for lognormal (left) and normal (centre) functions,
and discrete fragility curves (right), developed by Barbat et al. (2006), Ceran (2010) and
D’Ayala et al. (1997), respectively. ........................................................................ 33
Figure 18. Distribution of the examined methodologies over the years. ...................... 49
Figure 19. Distribution of the examined methodologies by country (left), building type
used in the analysis (centre), and structural material (right). ................................... 50
Figure 20. Distribution of the examined methodologies regarding capacity: components
for structural assessment (a); classes of building height (b); analysis type (c); model
type (d); shear failure (e); out-of-plane mechanism (f); horizontal diaphragms (g) and
geometrical irregularities (h). ............................................................................... 51

67
Figure 21. Distribution of the examined methodologies regarding demand:
representation of seismic demand (left), and consideration of site-specific seismic
demand input data (right). ................................................................................... 52
Figure 22. Distribution of the examined methodologies regarding the damage states
thresholds definition (a), IMs (b), sample size (c), and sampling method (d) .............. 53
Figure 23. Distribution of the examined methodologies regarding the uncertainty in
capacity (left), demand (centre), and damage state thresholds (right) ....................... 53

68
List of tables

Table 1. Main features of the considered methodologies for deriving analytical fragility
curves. .............................................................................................................. 37
Table 2. Qualitative assessment criteria and respective rating................................... 41
Table 3. Application of the qualitative assessment criteria on the reviewed fragility
curves. .............................................................................................................. 45

69
Europe Direct is a service to help you find answers to your questions about the European Union
Free phone number (*): 00 800 6 7 8 9 10 11
(*) Certain mobile telephone operators do not allow access to 00 800 numbers or these calls may be billed.

A great deal of additional information on the European Union is available on the Internet.
It can be accessed through the Europa server http://europa.eu

How to obtain EU publications

Our publications are available from EU Bookshop (http://publications.europa.eu/howto/index_en.htm),


where you can place an order with the sales agent of your choice.

The Publications Office has a worldwide network of sales agents.


You can obtain their contact details by sending a fax to (352) 29 29-42758.
LB-NA-27635-EN-N
JRC Mission

As the Commission’s
in-house science service,
the Joint Research Centre’s
mission is to provide EU
policies with independent,
evidence-based scientific
and technical support
throughout the whole
policy cycle.

Working in close
cooperation with policy
Directorates-General,
the JRC addresses key
societal challenges while
stimulating innovation
through developing
new methods, tools
and standards, and sharing
its know-how with
the Member States,
the scientific community
and international partners.

Serving society
Stimulating innovation
Supporting legislation

doi:10.2788/586263

ISBN 978-92-79-54136-0

You might also like