0% found this document useful (0 votes)
103 views156 pages

2024 - JRC-Guidance On The Design For Structural Robustness

Uploaded by

zdrico1
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
103 views156 pages

2024 - JRC-Guidance On The Design For Structural Robustness

Uploaded by

zdrico1
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 156

ISSN 1831-9424

Guidance on the design for structural


robustness

Support to the implementation,


harmonization and further development of the Eurocodes

Andre, J., Anghileri, M., Belletti, B., Biondini, F., Caspeele,


R., Demonceau, J., Izzuddin, B., Martinelli, P., Molkens, T.,
O'Connor, A., Parisi, F., Sio, J., Sousa, M.L., Thienpont, T.

Edited by: Caspeele, R., Thienpont, T., Sousa, M.L.

2024

EUR 32018
This document is a publication by the Joint Research Centre (JRC), the European Commission’s science and knowledge
service. It aims to provide evidence-based scientific support to the European policymaking process. The contents of this
publication do not necessarily reflect the position or opinion 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. For information on the methodology and quality underlying the data used in this publication for which the
source is neither Eurostat nor other Commission services, users should contact the referenced source. The designations
employed and the presentation of material on the maps do not imply the expression of any opinion whatsoever on the
part of the European Union concerning the legal status of any country, territory, city or area or of its authorities, or
concerning the delimitation of its frontiers or boundaries.

EU Science Hub
https://joint-research-centre.ec.europa.eu
JRC138689

EUR 32018

Print ISBN 978-92-68-20640-9 ISSN 1018-5593 doi:10.2760/4945167 KJ-NA-32-018-EN-C


PDF ISBN 978-92-68-19818-6 ISSN 1831-9424 doi:10.2760/525706 KJ-NA-32-018-EN-N

Luxembourg: Publications Office of the European Union, 2024

© European Union / Andre, J. / Anghileri, M. / Belletti, B. / Biondini, F. / Caspeele, R. / Demonceau, J. / Izzuddin, B. / Martinelli,
P. / Molkens, T. / O'Connor, A. / Parisi, F. / Sio, J. / Thienpont, T., 2024

The reuse policy of the European Commission documents is implemented by the Commission Decision 2011/833/EU of 12 December 2011
on the reuse of Commission documents (OJ L 330, 14.12.2011, p. 39). Unless otherwise noted, the reuse of this document is authorised
under the Creative Commons Attribution 4.0 International (CC BY 4.0) licence (https://creativecommons.org/licenses/by/4.0/). This means that
reuse is allowed provided appropriate credit is given and any changes are indicated.

For any use or reproduction of photos or other material that is not owned by the European Union permission must be sought directly from
the copyright holders.

How to cite this report: European Commission, Joint Research Centre, Andre, J., Anghileri, M., Belletti, B., Biondini, F., Caspeele, R., Demonceau,
J., Izzuddin, B., Martinelli, P., Molkens, T., O`Connor, A., Parisi, F., Sio, J., Sousa, M.L. and Thienpont, T., Guidance on the design for structural
robustness, Caspeele, R., Thienpont, T. and Sousa, M.L. editor(s), Publications Office of the European Union, Luxembourg, 2024,
https://data.europa.eu/doi/10.2760/525706, JRC138689.
Contents

Abstract ............................................................................................................... 1
Foreword .............................................................................................................. 2
Report Series “Support to the implementation, harmonization and further development of
the Eurocodes” ...................................................................................................... 5
Acknowledgements ................................................................................................ 7
List of authors and editors ...................................................................................... 7
1 Introduction and scope ....................................................................................... 9
1.1 Introduction ................................................................................................ 9
1.2 Scope ....................................................................................................... 10
1.3 Terms and definitions ................................................................................. 10
1.3.1 Context ............................................................................................ 10
1.3.2 Hazard ............................................................................................. 10
1.3.3 Hazardous scenario ............................................................................ 10
1.3.4 Event ............................................................................................... 11
1.3.5 Hazardous event ................................................................................ 11
1.3.6 Exposure .......................................................................................... 11
1.3.7 Load path ......................................................................................... 11
1.3.8 Damage............................................................................................ 11
1.3.9 Failure mode ..................................................................................... 11
1.3.10 Consequences ................................................................................... 11
1.3.11 Progressive collapse ........................................................................... 11
1.3.12 Disproportionate collapse .................................................................... 11
1.3.13 Fragility ............................................................................................ 11
1.3.14 Vulnerability ...................................................................................... 11
1.3.15 Damage tolerance .............................................................................. 11
1.3.16 Continuity ......................................................................................... 12
1.3.17 Ductility ............................................................................................ 12
1.3.18 Integrity ........................................................................................... 12
1.3.19 Uncertainties ..................................................................................... 12
1.3.20 Probability ........................................................................................ 12
1.3.21 Reliability.......................................................................................... 12
1.3.22 Structural safety ................................................................................ 12
1.3.23 Risk ................................................................................................. 12
1.3.24 Redundancy ...................................................................................... 12
1.3.25 Foreseen event .................................................................................. 12

i
1.3.26 Unforeseen event............................................................................... 12
References...................................................................................................... 12
2 Main principles on structural robustness ............................................................. 13
2.1 Concept of robustness and definitions .......................................................... 13
2.1.1 Introduction ...................................................................................... 13
2.1.2 Disproportionate and progressive collapse ............................................ 17
2.1.3 Structural robustness and collapse resistance ........................................ 18
2.1.4 Structural robustness, redundancy, and static indeterminacy .................. 19
2.1.5 Structural robustness, vulnerability, and structural integrity .................... 19
2.1.6 Requirements to measure structural robustness .................................... 20
2.1.7 Concept of robustness in various fields ................................................. 20
2.2 Hazards and damage scenarios ................................................................... 21
2.2.1 Type of hazards ................................................................................. 21
2.2.2 Continuous damage and life-cycle robustness ........................................ 23
2.2.3 Damage modelling ............................................................................. 24
2.2.4 Damage propagation .......................................................................... 25
2.2.5 Modelling of hazards .......................................................................... 25
2.3 Consequences of failure .............................................................................. 26
2.3.1 Direct and indirect consequences ......................................................... 26
2.3.2 Consequences in the Eurocodes ........................................................... 28
2.3.3 Factors affecting the consequences of failure ......................................... 29
References ......................................................................................................... 31
3 Overview of current standardization and design guidelines .................................... 35
3.1 Existing national standards, regulations and practices .................................... 35
3.1.1 Development and evolution of current design guidelines ......................... 35
3.1.2 Existing European provisions for robustness .......................................... 37
3.1.3 Provisions for robustness outside Europe .............................................. 47
3.2 Assessment of current design guidelines ....................................................... 53
3.2.1 Indirect Design methods ..................................................................... 53
3.2.2 Direct Design methods ....................................................................... 55
3.2.3 Systematic Risk assessment ................................................................ 57
3.3 Flaws and inconsistencies in current standardization ...................................... 57
References ......................................................................................................... 59
4 State-of-the-art design considerations, approaches and strategies for improving
robustness .......................................................................................................... 61
4.1 General .................................................................................................... 61
4.2 Design strategies to prevent disproportionate collapse ................................... 62
4.3 Design considerations ................................................................................ 62
4.3.1 Classification of accidental actions ....................................................... 63

ii
4.3.2 Consequence Classes ......................................................................... 63
4.4 Design approaches ..................................................................................... 65
4.4.1 Definition of design scenarios .............................................................. 65
4.4.2 Design for specified accidental actions .................................................. 66
4.4.3 Design for actions from unspecified threats ........................................... 66
4.5 Strategies for improving robustness ............................................................. 67
4.5.1 General ............................................................................................ 67
4.5.2 Alternative load path strategy ............................................................. 68
4.5.3 Consequence reduction strategy .......................................................... 68
4.5.4 Event control strategy ........................................................................ 69
4.5.5 Key element design strategy ............................................................... 69
4.5.6 Prescriptive rules ............................................................................... 70
4.6 Design for robustness against ageing and deterioration .................................. 70
4.7 Multi-hazard design considerations .............................................................. 71
4.8 Seismic design versus robustness ................................................................ 72
4.9 Fire design versus robustness ..................................................................... 75
References ......................................................................................................... 77
5 Quantitative measures of structural robustness ................................................... 81
5.1 Structural performance indicators for robustness assessment .......................... 81
5.1.1 Deterministic indicators ...................................................................... 82
5.1.2 Probability-based and reliability-based indicators ................................... 86
5.1.3 Risk-based indicators ......................................................................... 86
5.2 Structural analysis methods for robustness assessment .................................. 87
5.2.1 General ............................................................................................ 87
5.2.2 Numerical modelling strategies ............................................................ 88
5.2.3 Response of single components to abnormal loads ................................. 90
5.2.4 Linear and limit analysis methods for structural systems (Level I) ............ 93
5.2.5 Nonlinear analysis methods for structural systems (Level II) ................... 95
5.2.6 Probabilistic analysis methods for structural systems (Level III) .............. 95
5.3 Damage-based robustness quantification ...................................................... 97
References ......................................................................................................... 99
6 Novel proposals for robustness provisions ......................................................... 105
6.1 Building Structures .................................................................................. 105
6.1.1 Horizontal Tying Force Strategy ......................................................... 105
6.1.2 Application of different structural Systems/ Materials ........................... 108
6.1.3 Alternative load path strategy ........................................................... 115
6.1.4 Examples of the application of the novel proposals ............................... 119
6.2 Bridge Structures..................................................................................... 140

iii
References.................................................................................................... 142
List of figures .................................................................................................... 146
List of tables ..................................................................................................... 148

iv
Abstract
This JRC Technical Report presents scientific and technical background information to
introduce the different aspects involved in providing robustness of structures. It is intended
to bring together references and ongoing work on the subject as well as stimulate debate.
It presents background information, state-of-the-art references and discusses provisions
in available guidelines. As such, it serves as a basis for further work to achieve a
harmonized European view on the consideration of robustness in the design, execution and
assessment of structures. The report focusses on new-build construction, although the
underlying principles also apply to existing structures.
The report introduces the general principles of structural robustness, including concepts
and terminology, hazards and damage scenarios as well as assessment of the
consequences of failure. An overview is provided of current standardization and design
guidelines in Europe as well as outside of Europe. Strengths and weaknesses in current
provisions are discussed. State-of-the-art information is collected covering alternative
design strategies, approaches and considerations. Specific information on strategies to
improve robustness is outlined, including the importance of allowing for ageing and
deterioration, and aspects related to multi-hazard design. Whilst robustness as a design
principle covers a range of extreme design events, including seismic and fire, differences
in design approaches for such exposures are also important to recognize. State-of-the-art
research information is referenced where available. Finally, a series of novel proposals for
robustness provisions is provided encompassing more detailed technical guidance
concerning the tying force strategy, the alternative load path strategy, etc. are proposed
to encourage discussion.

1
Foreword
The construction ecosystem is of strategic importance to the European Union (EU), as it
delivers the buildings and infrastructures needed by the rest of the economy and society,
having a direct impact on the safety of persons and the quality of citizens’ life. The
construction ecosystem includes activities carried out during the whole lifecycle of buildings
and infrastructures, namely design, construction, maintenance, refurbishment and
demolition. The industrial construction ecosystem employs around 25 million people in the
EU and provides an added value of EUR 1 158 billion (9.6% of the EU total)1,2,3.
The construction ecosystem is a key element for the implementation of the European Single
Market and many other important EU strategies and initiatives. The European Green Deal
(COM(2019) 640 final) aims to achieve climate neutrality for Europe by 2050, and relies
on numerous initiatives, noteworthy:
— the New Circular Economy Action Plan (COM(2020) 98 final) and the New Industrial
Strategy for Europe (COM(2020) 102 final) intending to accelerate the transition of the
EU industry to a sustainable model based on the principles of circular economy;
— the revision (COM(2022) 144 final) of the Construction Products Regulation (Regulation
(EU) No 305/2011) aiming to enable the construction ecosystem’s contribution to
meeting climate and sustainability goals and embrace the digital transformation of the
built environment;
— the New EU Strategy on Adaptation to Climate Change (COM (2021) 82 final) supported
by the recent Commission Communication on managing climate risks (COM(2024) 91
final) that reinforces the need to address climate change concerns to guarantee the
resilience and sustainability of built structures and infrastructures and to ensure regular
science-based risk assessments;
— the first European Climate Risk Assessment (EUCRA) report which highlights the
importance of EU policies for the built environment, including updating construction
standards and related European datasets.
Furthermore and recognizing that the EU's ambitions towards a climate neutral, resilient
and circular economy cannot be delivered without leveraging the European standardization
system, the European Commission presented a new Standardization Strategy (COM(2022)
31 final). The strategy spots standards as “the silent foundation of the EU Single Market
and global competitiveness”.
The EU has put in place a comprehensive legislative and regulatory framework for the
construction sector, including European standards (EN). Within this framework, the
Eurocodes are a series of 10 European standards, EN 1990 to EN 1999, providing common
technical rules for the design of buildings and other civil engineering works. In fact, the
Commission Communication on managing climate risks directly mentions the Eurocodes,
highlighting the role of building and infrastructure standards in integrating climate
adaptation and resilience.
The Commission Recommendation 2003/887/EC on the implementation and use of the
Eurocodes for construction works and structural construction products recommends
undertaking research to facilitate the integration into the Eurocodes of the latest
developments in scientific and technological knowledge. In this context, the so-called
second generation of the Eurocodes is under development under Mandate M/515 and
expected to be available by 2026. The second generation Eurocodes incorporates
improvements to the existing standards and extends their scope by embracing new

1
Commission staff working document: Scenarios for a transition pathway for a resilient, greener and more digital
construction ecosystem (https://ec.europa.eu/docsroom/documents/47996)
2
Council of the EU, Press release 30 June 2023, https://www.consilium.europa.eu/en/press/press-
releases/2023/06/30/council-adopts-position-on-the-construction-products-regulation/
3
Transition Pathway For Construction, European Commission, DG GROW,
https://ec.europa.eu/docsroom/documents/53854

2
methods, new materials, and new regulatory and market requirements, including
considerations for climate change impact on structural design.
This background document is published as a part of the JRC Report Series “Support to the
implementation, harmonization and further development of the Eurocodes” and presents
background information, the state of the art and critical assessments relating to
technical guidance for the assessment of structural robustness in the design of
new structures. This document was developed by CEN/TC 250 Working Group
(WG) 6 on Robustness with the participation of the JRC. The purpose of this working
group is to bring together the different national views and approaches concerning structural
robustness and to develop a broadly accepted and coherent set of harmonised European
technical rules for structural robustness.
This JRC Report presents scientific and technical background intended to provide a clear
view of the different aspects involved when considering structural robustness and to
stimulate debate. As such, it serves as a basis for further work to achieve a harmonized
European view on the treatment of structural robustness.
We hope that this report will provide a sound and helpful basis for discussions about the
topic of structural robustness, in particular concerning standardization work.
The report is available to download from the “Eurocodes: Building the future” website
(http://eurocodes.jrc.ec.europa.eu).
The authors have sought to present useful and consistent information in this report.
However, users of the information contained in this report must assess if such information
is suitable for their purposes.

Ispra, 2024

François Augendre and Georgios Tsionis


Built Environment Unit
Directorate E – Societal Resilience and Security
Joint Research Centre (JRC)
European Commission

Julie Bregulla
The Engineering and Design Institute (TEDI), London, UK
Convenor CEN/TC 250/ WG 6 Robustness

Robby Caspeele
Ghent University, Belgium
Coordinator writing panel of the JRC Report

3
4
Report Series “Support to the implementation, harmonization
and further development of the Eurocodes”

In the light of the Commission Recommendation of 11 December 2003, DG JRC is


collaborating with DG GROW and CEN/TC 250 “Structural Eurocodes”, and is publishing the
Report Series “Support to the implementation, harmonization and further development of
the Eurocodes”. This Report Series includes, at present, the following types of reports:

1. Science for policy documents, conveying the implications of scientific and


technical evidence for a policymaking process;
2. Technical documents, facilitating the implementation and use of the Eurocodes
and containing information and practical examples (Worked Examples) on the use
of the Eurocodes and covering the design of structures or its parts (e.g. the technical
reports containing the practical examples presented in the workshop on the
Eurocodes with worked examples organized by the JRC);
3. Pre-normative documents, resulting from the works of the CEN/TC 250 and
containing background information and/or the first draft of proposed normative
parts. These documents can be then converted to CEN technical specifications;
4. Background documents, providing approved background information on the
current Eurocode part. The publication of the document is at the request of the
relevant CEN/TC 250 Sub-Committee;
5. Scientific/Technical information documents, containing additional,
non-contradictory information on the current Eurocode part, which may facilitate its
implementation and use, or preliminary results from pre-normative work and other
studies, which may be used in future revisions and further developments of the
standards. The authors are various stakeholders involved in the Eurocodes process
and the publication of these documents is authorized by the relevant CEN/TC 250
Sub-Committee or Working Group.

Editorial work for this Report Series is performed by the JRC together with partners
and stakeholders, when appropriate. The publication of the reports types 3, 4 and 5 is
made after approval for publication by CEN/TC 250, or the relevant Sub-Committee or
Working Group.

The publication of these reports by the JRC serves the purpose of implementation, further
harmonization and development of the Eurocodes. However, it is noted that neither the
Commission nor CEN are obliged to follow or endorse any recommendation or result
included in these reports in the European legislation or standardisation processes.

The reports are available to download from the website “Eurocodes: Building the future”
(http://eurocodes.jrc.ec.europa.eu).

5
6
Acknowledgements
This report has been prepared to support the development of future European guidance
for the assessment of structural robustness under the shield of CEN/TC 250. Both CEN/TC
250 and JRC acknowledge the substantial contribution of the many international experts
of CEN/TC 250/WG 6 and others, who have supported the works with their essential input
and reviews. In particular, the tremendous work from the writing panel of this JRC report
is recognized. Finally, the contribution of Artur Pinto and Silvia Dimova, formerly at JRC
Unit Safety & Security of Buildings (now Built Environment), is gratefully acknowledged.

List of authors and editors


Authors

João ANDRÉ Portuguese National Laboratory for Civil


Engineering, Portugal

Mattia ANGHILERI Politecnico di Milano, Italy

Beatrice BELLETTI University of Parma, Italy

Fabio BIONDINI Politecnico di Milano, Italy

Robby CASPEELE Ghent University, Belgium

Jean-François DEMONCEAU University of Liège, Belgium

Bassam IZZUDDIN Imperial College London, United Kingdom

Paolo MARTINELLI Politecnico di Milano, Italy

Tom MOLKENS KU Leuven, Belgium

Alan O’CONNOR Trinity College Dublin, Ireland

Fulvio PARISI University of Naples Federico II, Italy

João SIO Imperial College London, United Kingdom

Maria Luisa SOUSA Portuguese National Laboratory for Civil


Engineering, Portugal, formerly Joint
Research Centre (JRC), European
Commission

Thomas THIENPONT Ghent University, Belgium

Editors

Robby CASPEELE Ghent University, Belgium

Thomas THIENPONT Ghent University, Belgium

Maria Luisa SOUSA Portuguese National Laboratory for Civil


Engineering, Portugal, formerly Joint
Research Centre (JRC), European
Commission

7
8
1 Introduction and scope

1.1 Introduction
Structural failures, especially involving progressive collapse, such as Ronan Point, tragically
bring to light the importance of robustness in design – both to the structural engineering
community and society at large. Failures can be caused by extreme or exceptional loading
and damaging events, though also under standard, to be expected, loading scenarios and
structural deterioration processes, and can often be deemed disproportionate. National
regulations across Europe and the world and accompanying design standards outline
requirements for buildings to be designed to resist disproportionate failure. These
provisions generally require structural elements to be tied together, adding redundant
structural members, which have sufficient strength and stiffness to resist, etc.
Over the past decades, some accidental or deliberate exceptional events (such as
explosions, impact, unforeseen material degradation, etc.) reveal the vulnerability that
structures can have to localized damage. The importance of buildings to resist progressive
collapse remains one of the contemporary design aims of the structural engineering
community. It remains a developing field, with most research undertaken over the last 40
years, and as such brings a need to regularly review and adjust codes of practice and
guidelines to keep pace with knowledge.
This JRC Report presents scientific and technical background information to introduce the
different aspects involved in providing robustness to structures. It is intended to bring
together references and ongoing work on the subject as well as stimulate debate. It
presents background information, state-of-the-art references and discusses provisions in
available guidelines. As such, it serves as a basis for further work to achieve a harmonized
European view on the consideration of robustness in the design, execution and assessment
of structures. The report focuses on new-build construction, although the underlying
principles also apply to existing structures.
The report consists of six main chapters.
● Chapter 1: General introduction, scope and relevant terms and definitions
● Chapter 2: General principles related to structural robustness.
● Chapter 3: An overview is provided of current standardization and design
guidelines in Europe and outside Europe, together with some background
information, when available. Strengths and weaknesses in current provisions
are discussed.
● Chapter 4: New state-of-the-art information is collected covering alternative
design strategies, approaches and considerations. Specific information on
strategies to improve robustness is outlined, including the importance of
allowing for ageing and deterioration and aspects related to multi-hazard
design. Whilst robustness as a design principle covers a range of extreme design
events, including seismic and fire, differences in design approaches for such
exposures are also important to recognize.
● Chapter 5: Based on more recent developments, state-of-the-art research
information is provided concerning quantitative measures and calculation
methods for robustness. An overview of performance indicators for structural
robustness is provided as well as an overview of the structural analysis methods
with their abilities and limitations.
● Chapter 6: A series of novel proposals for robustness provisions is provided
based on the work performed by the M/515 mandated work of project team 2
of CEN/TC250/WG6 encompassing more detailed technical guidance concerning
the tying force strategy, the alternative load path strategy, etc.

9
1.2 Scope
The topic of structural robustness is complex and subject to National regulatory provisions,
across Europe. In fact, every country has specific regulations about robustness of
structures. There are many local practices throughout Europe to achieve robustness. The
complexity and diversity of the topic are illustrated by these approaches, ranging from
conceptual principles to be considered in the design phase, to prescriptive rules and
quantitative design measures that are available in specialist literature. Furthermore,
robustness strategies and provisions differ depending on the materials and structural
typologies and forms used. As such, finding a harmonized view is a difficult challenge for
the standardization work in the coming years. The report provides a broad view of the
topic, definitions and practices as well as universal general principles. This work brings
together key information to enable the debate that will support a common view as the
standardization for robustness is progressed.
The information contained in the report is intended for a broad range of users,
encompassing designers, decision-makers, third-party control organizations, experts, etc.
As such, not all parts of the report might be of equal importance to a particular user.
Depending on the needs of several users, the reader is provided with references and
supporting information, aligned with complementary and more detailed provisions where
relevant.
Although the authors have attempted to provide a broad overview of the available state-of-
the-art knowledge in the field, the intention has not been to provide a complete catalogue,
but rather to provide sufficient information to understand the different aspects involved
and guide readers to more detailed information if they wish to learn more about the topic.
In particular, relevant background information was collected and assistance was given in
relation to some of the principles underlying the Eurocode approach to robustness. Further,
potentially relevant information that could be considered for the further evolution of the
Eurocode in relation to robustness is highlighted. However, it should be acknowledged that
the topic of structural robustness is a fast-paced and evolving research field where research
is still ongoing; this is highlighted in the report.
Finally, it is important to note that the information presented here is background
information for users seeking a more thorough understanding and appreciation of ongoing
work in this field. First and foremost designers must consult the Eurocodes and relevant
National provisions for their designs and stakeholder interactions. It is hoped that the
compilation of this material will support those addressing this most important design
consideration and allow them to arrive at informed decisions, enabling them to gain an
efficient and swift overview of relevant references.

1.3 Terms and definitions

1.3.1 Context
Set of all relevant circumstances within which engineering decisions are made.

1.3.2 Hazard
Exceptionally unusual and severe threat, e.g. a possible abnormal action or environmental
influence, insufficient strength or stiffness, or excessive detrimental deviation from
intended dimensions (ISO 2394, 2015).

1.3.3 Hazardous scenario


Series of situations, transient in time, that a system might happen to undergo, and which
may endanger the system itself, people, and the environment (ISO 2394, 2015).

10
1.3.4 Event
Occurrence, or change, of a particular set of circumstances.

1.3.5 Hazardous event


Occurrence of a hazard or a hazardous scenario.

1.3.6 Exposure
Set of different events that could act on the constituents of the system with potential
consequences for the considered system.

1.3.7 Load path


Integral of all elements of the system affected by action effects, from the point of
application of the action to the boundaries of the system.

1.3.8 Damage
Unfavourable change in the condition of a system that can affect the performance of the
latter.

1.3.9 Failure mode


Description of how failures propagate within the structural system.

1.3.10 Consequences
Adverse outcome of a hazardous event. Consequences can be defined to be restricted to
the performance of the structural system or as having a wide scope, e.g. including human,
economic and environment qualities.

1.3.11 Progressive collapse


Characteristic of a failure mode where an initially localised failure event leads to a sequence
of follow-up failure events in a domino style.

1.3.12 Disproportionate collapse


Characteristic of a failure mode where there is a distinct disproportion the immediate
damage following a triggering event and the follow-up failure events.
NOTE: It would be desirable that a discrimination between proportionate and
disproportionate collapse could be made on the basis of a range of area or percentage of
members failed, but such quantitative proposal are currently not readily available in
literature or standardization documents.

1.3.13 Fragility
System characteristic expressing its structural performance, typically in terms of
probability of exceedance of a certain limit state conditional to the occurrence of a specific
intensity measure.

1.3.14 Vulnerability
Describes the degree of susceptibility of a structural system to attain a particular level of
consequences, for a given hazardous event.

1.3.15 Damage tolerance


Ability of a structural system to sustain a given level of damage while maintaining
equilibrium with the applied loads.

11
1.3.16 Continuity
Continuous connection of members of a structural system.

1.3.17 Ductility
Ability of a structural system to sustain the applied loads by dissipating plastic energy.

1.3.18 Integrity
Condition of a structural system to enable force transfer among members in case of
accidental events.

1.3.19 Uncertainties
State of deficient information, e.g. related to the understanding, or knowledge of, an event,
its consequence, or likelihood.

1.3.20 Probability
Mathematical expression of the degree of confidence in a prediction.

1.3.21 Reliability
Probabilistic measure of the ability of a structural system to fulfil specific design
requirements. Reliability is commonly expressed as the complement of the probability of
failure.

1.3.22 Structural safety


Quality of a structural system, referring to the strength, stability and integrity of a structure
to withstand the hazards to which it is likely to be exposed during its life-time.

1.3.23 Risk
A measure of the combination (usually the product) of the probability or frequency of
occurrence of a defined hazard and the magnitude of the consequences of the occurrence.

1.3.24 Redundancy
The ability of the system to redistribute among its members the load which can no longer
be sustained by some damaged and/or deteriorated elements

1.3.25 Foreseen event


An event to which the structure can be subjected and of which the nature, the magnitude,
and the probability of occurrence during the construction or use of the structure, can be
defined.

1.3.26 Unforeseen event


An event to which the structure can be subjected, but which occurrence cannot be defined,
until it materializes.

References
ISO. (2015). ISO2394 (Fourth edition) - General principles on reliability for structures.
International standardization Organization.

12
2 Main principles on structural robustness

2.1 Concept of robustness and definitions

2.1.1 Introduction
During the last few decades, increasing attention has been focused on the concepts of
structural robustness, disproportionate failure and progressive collapse. This is due to the
continued occurrence of cases in which the structures are involved in disproportionate
collapse with respect to an initial localized structural damage, generated by an accidental
action, like explosions and impacts, or underestimated events, like deterioration processes.
It is noted that the field of robustness only deals with unforeseen events and loading. Other
accidental events are being dealt with by application of Eurocodes.
The first iconic failure in this field was the partial collapse in 1968 of the Ronan Point
building in London after a relatively small gas explosion led to the progressive failure of its
entire corner (Figure 2-1). The building was a 22-storey residential building, constructed
in 1966, assembling precast concrete panels together without a structural frame system.
Many connections relied on friction only and each floor was supported by the load-bearing
walls directly beneath it. In May 1968, a gas explosion on the 18th floor caused the failure
of an external precast concrete panel creating a progressive collapse (chain reaction)
upward and then downward due to dynamic effects until the entire southeast corner of the
structure collapsed. Moreover, when the building was dismantled, signs of poor
workmanship was found (Pearson & Delatte 2005).
The collapse of the Ronan Point building may be attributed to the lack of structural
redundancy and robustness of the system (no alternative load paths could be activated
following the removal of a single component). A relatively small initial damage ended with
a disproportionate effect, namely the collapse of the building corner.
Another failure event which increased the attention of the concept of disproportionate
collapse and structural robustness concerns the Alfred Murrah Federal Building (Oklahoma
City, Oklahoma, USA). The federal structure was a 9-storey reinforced concrete frame
building with shear walls. The stability of the system was governed by a transfer girder at
the third-floor level and the lower storey columns supporting it (Corley et al. 1998). In
1995, a truck bomb explosion outside the building resulted in the failure of three main
columns supporting the transfer girder. Almost half of the building collapsed (Figure 2-2).
Records indicated an extremely well designed and detailed structure. The terrorist attack
resulted in the loss of three lower columns and a portion of the surrounding floors as direct
consequences of the explosion. But a progressive collapse extended the damage
disproportionately, far beyond the structural failure directly caused by the bomb explosion.
Other building collapse events, including the attacks to the twin towers of the World Trade
Center and Pentagon Building in 2001, emphasize the need for additional considerations in
structural design codes regarding the concept of disproportionate collapse and robust
structures (Carper & Smilowitz 2006). The Pentagon Building is a five-storey reinforced
concrete building subdivided into five segmented concentric rings separated by expansion
joints (Figure 2-3). During the terrorist attack in September 2001, an aircraft hit the
building and the damage involved only the impact area thanks to the structural
compartmentalisation of the building. Similar results occurred to the Charles de Gaulle
Airport (Paris, France). In 2004, a portion of the concrete shell roof of the airport collapsed
(Figure 2-4).
Progressive and disproportionate failures are not limited to buildings. They are also
recognised as a major concern for the collapse of bridges (Starossek 2008, National
Transportation Safety Board 1984). An example is the progressive failure of the Heang-Ju
Bridge (Seoul, South Korea), a continuous prestressed concrete girder bridge. In 1992,
after the failure of a temporary pier in the main span, the damage propagated through ten
adjacent spans (Figure 2-5). To this event the continuous prestressing tendons in the deck
played a significant and disastrous role by creating a chain between the spans. The opposite

13
situation occurred in 1975, when the discontinuity of the prestressing tendons between
adjacent spans of the Tasman Bridge (Hobart, Australia) avoided a progressive collapse
(chain reaction) when a bulk carrier collided against two pylons. The two closest spans to
the impact area were the only damaged structural elements (Figure 2-5). More recent
bridge collapse events, like the I-35W Mississippi River Bridge failure (Mississippi,
Minnesota, USA) in 2007 and the Morandi Bridge event (Genoa, Italy) in 2018, increased
the attention on bridge design, execution and maintenance regarding disproportionate
collapse, structural integrity and robustness. The Morandi bridge event also highlights the
importance of considering the importance of specific robustness provision for isostatic
structures.

Figure 2-1. Collapse of Ronan Point Building, London, UK (1968) 4

Figure 2-2. Collapse of the Alfred Murrah Federal Building, Oklahoma City, USA (1995) 5

4
Source: https://www.architecture.com/explore-architecture/inside-the-riba-collections/people-in-high-places,
https://failedarchitecture.com/the-downfall-of-british-modernist-architecture/ and
https://failedarchitecture.com/the-downfall-of-british-modernist-architecture/
5
Source: https://en.wikipedia.org/wiki/Alfred_P._Murrah_Federal_Building and
https://www.kpbs.org/news/arts-culture/2017/02/02/american-experience-oklahoma-city

14
Figure 2-3. Terrorist attack - The Pentagon Building, Washington D.C., USA (2001) 6

Figure 2-4. Collapse of Terminal 2E of the Charles de Gaulle Airport, Paris, France (2004) 7

Figure 2-5. Collapse of (a) Heang-Ju Bridge, Seoul, South Korea (1992) and (b) and Tasman
Bridge, Hobart, Australia (1975) 8
(a) (b)

Optimisation techniques in contemporary design and the and the speed of the execution
phase without adequate quality control may lead to unsatisfactory levels of structural
redundancy and robustness, thereby increasing the risk of disproportionate collapse when
accidental or unforeseen events occur.

6
Source: https://wjla.com/news/september-11th-20th-anniversary/gallery/how-the-pentagon-building-saved-
lives-on-911-terrorist-attacks-collapse-crashing-september-11?photo=10 and
https://en.wikipedia.org/wiki/File:Aerial_view_of_the_Pentagon_during_rescue_operations_post-
September_11_attack.JPEG
7
Source: https://engineers-channel.blogspot.com/p/charles-de-gaulle-airport-terminal-2e.html
8
Source: https://en.wikipedia.org/wiki/Haengju_Bridge and
https://en.wikipedia.org/wiki/Tasman_Bridge_disaster

15
In structural design codes and standards, the term robustness has been initially used to
indicate the ability of a system to resist damage under extreme loads. Afterwards, since
the early 1980s, British codes have included requirements to prevent the progressive
spreading of an initial local failure leading to a disproportionate collapse.
However, in structural design, the concept of robust structures is still an issue of
controversy. In fact, despite the fact that procedures aimed to identify weak links within
structures have been reported in literature (Lu et al. 1999, Agarwal et al. 2003) and efforts
have been made either to propose design strategies to prevent progressive and
disproportionate collapse (Starossek & Wolff 2005, Ellingwood & Dusenberry 2005,
Haberland & Starossek 2010, André & Faber 2019) or quantify robustness (Baker et al.
2008, André et al. 2015), there are no well-established and generally accepted criteria for
a consistent definition and a quantitative measure of structural robustness (Starossek &
Haberland 2011, André 2020). Recently, advances in robustness quantification have been
accomplished for deteriorating structural systems (Biondini & Restelli 2008, Biondini 2009,
Biondini & Frangopol 2014).
An important issue related to structural robustness is its perception, leading to significant
differences in its definition and quantification in structural codes and scientific publications.
Moreover, design strategies to avoid progressive collapse in literature and partially covered
by design standards do not, in general, distinguish between the concepts of collapse
resistance, structural redundancy, and structural robustness.
A selection of structural robustness definitions proposed in the last decades is reported in
Table 2-1.

Table 2-1. Definitions related to the concept of structural robustness

Reference Structural Robustness Definition

GSA, 2003 The ability of a structure or structural components to resist damage


without premature and/or brittle failure due to events like explosions,
impacts, fire or consequences of human error, due to its vigorous
strength and toughness.

Starossek & The term robustness is defined as insensitivity to local failure, with
Haberland, “insensitivity” and “local failure” being quantified by the applicable
2005 design criteria. Robustness is a property of the structure alone and
independent of the loading.

Val & Val, 2006 The ability of a structure to absorb the effect of an accidental event
without suffering damage disproportionate to the event that caused
it. […] ability of the structure to withstand local damage without
disproportionate collapse.

CEN - Eurocode The ability of a structure to withstand events like fire, explosions,
1. Part 1-7, impact or the consequences of human error without being damaged
2006 to an extent disproportionate to the original cause.

Bontempi et al., The robustness of a structure, intended as its ability not to suffer
2007 disproportionate damage as a result of limited initial failure, is an
intrinsic requirement, inherent to the structural system organization.

Vrouwenvelder, The notion of robustness is that a structure should not be too sensitive
2008 to local damage, whatever the source of damage […].

16
Agarwal & Robustness is the ability of a structure to avoid disproportionate
England, 2008 consequences in relation to the initial damage.

Biondini & Structural robustness can be viewed as the ability of the system to
Restelli 2008 suffer an amount of damage not disproportionate with respect to the
causes of the damage itself.

Narasimhan & A structure shall not be damaged by events like fire, explosions or
Faber, 2009 consequences of human errors, deterioration effects etc. to an extent
disproportionate to the severeness of the triggering event.

Gulvanessian et Property of structures that enables them to withstand unforeseen or


al., 2012 unusual circumstances without unacceptable levels of consequences
or intolerable risks.

fib Model Code, Robustness is a specific aspect of structural safety that refers to the
2013 ability of a system subject to accidental or exceptional loadings (such
as fire, explosions, impact or consequences of human errors) to
sustain local damage to some structural components without
experiencing a disproportionate degree of overall distress or collapse.

André et al., Measure of the predisposition of a structural system to loss of global


2015 equilibrium, as a result of a failure event for a given hazard scenario.

Considering these definitions, structural robustness is most often associated with the ability
of the system to avoid structural consequences that are disproportionate with respect to
the extent of the triggering initial damage.
In general, robustness evaluations should not be restricted to cases involving only
accidental actions, like explosions or impacts. Other cases are relevant, for example
involving the continuous damage over the structures lifetime. The effects of ageing and
deterioration process on civil engineering structures can lead over time to unsatisfactory
structural performances disproportionate with respect to the corresponding damage (Zhu
& Frangopol 2012). Although EN1990 stipulates that structures should be designed,
executed and maintained adequately by qualified and experienced people and controlled
by appropriate quality management, the occurrence of errors and negligence during the
design, execution and operation phases (which could lead to significant and disastrous
consequences) can never be fully excluded and for such situations robustness provisions
can limit the consequences resulting from such errors or negligence.
However, for most structures, design in accordance with the Eurocodes is assumed to
provide an adequate level of robustness without the need for any additional design
measures to enhance structural robustness.9

2.1.2 Disproportionate and progressive collapse


Disproportionate and progressive collapse denote two different aspects of the evolution of
structural damage: the concept of disproportionality between cause and effect and the
concept of failure progression.
Disproportionate collapse refers to a significant disproportion between the initial damage
event and the ensuing consequences, for example following the collapse of a major part or
even the whole of a structure. This is simply a subjective judgment made about
observations of the consequences following the occurrence of an initial damage event.
Instead, progressive collapse refers to the collapse evolution, describing how, after a local

9
Formulation in accordance to Note 2 in Clause 4.4(1) in FprEN1990:2022

17
initial event, failure may propagate to members other than the ones directly affected by
the initial damage, leading to a chain reaction between elements. A collapse may be
progressive horizontally (e.g. failure in succession of adjacent structural bays) or
progressive vertically (e.g. failure in succession of columns supporting a certain number of
floors) or a combination of both.
In general, a progressive collapse can result in a disproportionate collapse if the successive
failures involve a large part of the structure with respect to the triggering initial event. On
the other hand, a disproportionate collapse can be immediate or progressive. Therefore, a
collapse may be progressive but not necessarily disproportionate in its extents, for instance
if the damage is arrested after the propagation through a limited number of structural
bays. On the other hand, a collapse may be disproportionate but not necessarily
progressive. A selection of disproportionate and progressive collapse definitions proposed
in literature is listed in Table 2-2 and Table 2-3, respectively.
The concept of structural robustness is related to disproportionate effects of damage with
respect to the initial causes or amount of the damage itself. However, the relationship
between initial and final damage, which may be different due to a disproportionate and/or
progressive damage process, is not sufficient to quantify structural robustness. The amount
of damage has to be compared with the corresponding consequences. If damage does not
lead to disproportionate consequences, the system is considered robust.

Table 2-2. Definitions related to the concept of disproportionate collapse

Reference Disproportionate Collapse Definition

Agarwal & Disproportionate collapse results from small damage or a minor action
England, 2008 leading to the collapse of a relatively large part of the structure.

Starossek & A collapse that is characterized by a pronounced disproportion


Haberland, between a relatively minor event and the ensuing collapse of a major
2010 part or the whole of a structure.

Table 2-3. Definitions related to the concept of progressive collapse

Reference Progressive Collapse Definition

ASCE, 2016 Progressive collapse is defined as the spread of an initial local failure
from element to element resulting, eventually, in the collapse of an
entire structure or a disproportionately large part of it.

Ellingwood, A progressive collapse initiates as a result of local structural damage


2006 and develops, in a chain reaction mechanism, into a failure that is
disproportionate to the initiating local damage.

Canisius et al., Progressive collapse, where the initial failure of one or more
2007 components results in a series of subsequent failures of components
not directly affected by the original action is a mode of failure that can
give rise to disproportionate failure.

2.1.3 Structural robustness and collapse resistance


Structural robustness of a damaged system is a concept useful to prevent disproportionate
structural collapse. Robustness refers to the ability of a structure to resist initiating events
without exhibiting disproportionate performance reduction with respect to the causes or
amount of damage. Generally, in a robust structure, no failures disproportionate to the

18
initial damage will occur. Instead, the concept of collapse resistance can be regarded as
the insensitivity of a structure to abnormal collapse events (Haberland & Starossek 2009,
André et al. 2015).
For practical assessment, in case one wants to verify the ability of the structure to
withstand the initiating event without the need to quantify and define the maximum extent
of the progressive damage, a simplified and conservative verification of structural
robustness could take basis in the verification of equilibrium of the considered part of the
structure after the damage associated to the initiating event.

2.1.4 Structural robustness, redundancy, and static indeterminacy


The terms structural robustness, redundancy and static indeterminacy are often used as
synonymous. However, they denote different properties of the structural system.
Robustness is related to the ability of the system to suffer a loss of performance not
disproportionate with respect to the causes or amount of damage. Structural redundancy
can instead be defined as the ability of the system to redistribute among its member the
load which can no longer be sustained by other damaged and/or deteriorated elements. It
may be affected by several factors, such as structural topology, member sizes, material
properties, applied load and load sequences, among others (Frangopol & Curley 1987,
Frangopol & Klisinski 1989, Frangopol & Nakib 1991, Ghosn et al. 2010).
The additional effort affecting the remaining system components after a local failure may
reduce the structural performance and therefore lead to an unsatisfactory robustness level.
Damage propagation mechanisms under ageing and structural deterioration may also
involve disproportionate effects and alternative load redistribution paths, altering structural
redundancy and structural robustness (Okasha & Frangopol 2010, Biondini & Frangopol
2017). Alternative load paths may enhance structural redundancy. However, increasing
structural redundancy does not necessarily lead to increase structural robustness (Biondini
2009, André 2020). In fact, the redistribution of internal actions on the damaged system
may promote the evolution of damage reducing robustness. The collapse of the Nanfang’ao
Bridge (Taiwan) in 2019, may be related to lack of structural robustness. The 21-year-old
tied-arch steel bridge failed after the collapse of a single supporting cable. The
redistribution of internal forces after the initial damage resulted in a progressive cables
collapse.
Redundancy is usually associated with the degree of static indeterminacy. However, the
degree of static indeterminacy is not a consistent measure for structural redundancy and
increasing the degree of static indeterminacy does not necessarily lead to increase
structural redundancy. Examples can be found in Biondini et al. (2008).

2.1.5 Structural robustness, vulnerability, and structural integrity


The term “structural vulnerability” is often used in literature to express different concepts.
Vulnerability can be adopted to describe the sensitivity of the performance of a structure
to damage events. Therefore, in this case, the term “vulnerability” is used to express the
susceptibility of a component (or a system) to some external action. In this approach,
vulnerability is a property of the system since it examines the effect of potential hazards
triggered by unexpected events or unforeseen load scenarios. For instance, an earthquake
happening in a site “A” may produce very different consequences from the same
earthquake event occurring in a different site “B”. The infrastructure in place “B” may be
more vulnerable because of its form, technical design, execution and condition properties.
A structure can be made collapse resistant by reducing the vulnerability of its members by
protection techniques or increasing the resistance of the structural elements.
Sometimes, the term ”vulnerability” has also been adopted in terms of the sensitivity of
damage leading to disproportionate consequences (Agarwal & Blockley 2007). In this
approach, a structure is vulnerable if small damages lead to disproportionate failures.

19
Within this definition it is clear how vulnerability is considered as antagonistic to structural
robustness.
The broad distinction however is related to consider the vulnerability as the conditional
potential damage to a system or a state condition as a result of a hazard event. Therefore,
in the first case, vulnerability is simply the susceptibility to damage; in the second case, it
expresses the idea of the susceptibility in the sense that a small amount of damage can
lead to disproportionate consequences. However, the damage-sensitivity definition is
recommended to have a clear distinction between structural robustness and vulnerability
of a system.
Moreover, damage-sensitivity may also be associated with different levels of structural
integrity, that are associated with the severity of a potential structural failure with respect
to its consequences. In fact, the global collapse of a structural system is considered more
important than the local collapse of a single member or a portion of the system.

2.1.6 Requirements to measure structural robustness


In order to measure the amount of robustness of a structural system, some requirements
must be satisfied:
— The system and the design objectives and requirements must be clearly defined,
including robustness criteria and measures;
— The intended limit state functions of a system have to be identified and formulated,
since structural robustness provides relevant information with respect to certain
desirable system performance objectives;
— The hazardous events which may affect the structure have to be identified and the
corresponding structural damages have to be computed, at both the element and
system levels;
— The overall structural consequences of damaging scenarios have to be analysed with
regard to the mentioned limit state functions and compared with the corresponding
amount of damage. Two basic ingredients are hence necessary to be compared for
robustness evaluation: the global damage affecting the system and the corresponding
structural consequences;
— Proper consideration of the uncertainties affecting several input parameters, as a
function of the level of accuracy required.
Therefore, robustness evaluation is associated with specific structural performance
indicators, limit state conditions, and damage scenarios.

2.1.7 Concept of robustness in various fields


In general terms, without focusing on a specific scientific field, robustness is related to the
sensitivity of a qualitative or quantitative features of a system with respect to changes in
system state due to unexpected disturbances.
Robustness is therefore associated with certain performance characteristics influenced by
some perturbations on the system. In order to define a robust system it is fundamental
that the clarification of both the considered system performance and system perturbation
of interest be provided. The comparison between these two features of the system is
necessary for robustness evaluation.
Robustness interpretations are used in engineering as well as in other fields such as
biology, statistics, control theory, among others. A selection of definitions of robustness in
different scientific areas is reported in the following list:
— Structural Engineering: ability/property of a system to avoid a structural performance
variation (system performance) disproportionately larger with respect to the
corresponding damage (system perturbation).

20
— Software Engineering: ability of the software (system performance) to react
appropriately to abnormal circumstances (system perturbation), i.e. circumstances
outside of specifications (Meyer 1997).
— Statistics: a robust statistical technique is insensitive against small deviations in the
assumptions (system perturbation) (Huber 1996).
— Ecosystems: ability of a system (system performance) to maintain functions even with
changes in internal structure or external environment (system perturbation) (Callaway
et al. 2000).
— Design optimization: a robust solution of an optimization problem under uncertainty is
the one that has the best performance (system performance) against the worst
contingency that may arise (system perturbation) (Kouvelis and Yu 1997).

2.2 Hazards and damage scenarios


In the field of structural robustness, a hazard is identified as a serious threat to the integrity
of a structure and the safety of people (Canisius 2011). Instead, an exposure scenario is
a set of hazards that possibly affect the structural response during construction and
lifetime.
The characteristics of some hazardous events (such as the exact time of occurrence, its
intensity and spatial distribution) are unknown to a designer and therefore design codes
typically specify rules concerning unidentified actions, ranging from prescriptive rules (e.g.
tying of elements or protecting vulnerable components) to notional hazardous scenarios to
which a structure should be designed against (e.g. notional removal of element(s)). In
addition, design codes (e.g. Eurocode) account for explicit design procedures and
requirements for the identified accidental actions such as earthquake, fire, impact and
internal gas explosions.

2.2.1 Type of hazards


The possible hazards that may play a role in the safety assessment of structural systems
can be classified based on the nature of the event itself (Canisius 2011). Three categories
are identified:
1. Type of hazardous events given by natural events and unintentional anthropogenic
hazards, e.g. earthquake and wind actions. They can be combined with time-dependent
effects from ageing and deterioration processes.
2. Security related events, such as vandalism and malicious attacks, intentionally man-
made. Due to their high unpredictably, it is in general, convenient to adopt appropriate
measures to prevent their occurrence and limit the possible damage propagation over
the entire system, instead of designing a structure to be able to support these actions
(which may lead to very strictly requirements).
3. Human errors and negligence during the design, execution and operation phases.
Progressive increasing attention has been focused on this type of hazard after the iconic
failure event of the Ronan Point Building, 1968 (Ellingwood 1987). An efficient strategy
is to implement quality control procedures during the entire life of the structures, from
the preliminary design process to the demolition phase. Even simple periodic visual
inspections may significantly increase the knowledge of the structural condition over
time and can be used to plan appropriate maintenance and repair interventions.
Some (non-limitative) hazards corresponding to these three categories are:
1. Gas explosion; dust explosion; fire; impact by vehicles, aircrafts, ships …;
overloading; earthquake; landslide; mining subsidence, tornado or
typhoon/hurricanes/cyclone; avalanche; rockfall; high groundwater; flood; storm
surge; volcanic eruption; environmental attack; tsunami; …
2. Bomb explosion; fire; vandalism; …

21
3. Design or assessment error; material error; construction error; user error; lack of
maintenance; …
Moreover, in particular cases, some hazards are a direct consequence of previous hazard,
i.e. cascading effects, leading to more serious effects. For example:
— Tsunami after an earthquake;
— Fire after an earthquake;
— Fire following a gas explosion or bomb blast;
— Accelerated material deterioration following damage from accidental action on the
structure.
According to Bontempi et al. (2007), hazards can be divided into physical and logical
threats, as shown in Figure 2-6. Physical threats are all the possible hazards that may
create a damage or failure to the structure, further sub-divided into external hazards (e.g.
extreme accidental and environmental actions) and intrinsic hazards that include all the
undetected defects of the structure. Error threats include design, execution and operation
errors. Another critical hazard not always considered, that can be classified in the logical
group, is lack of maintenance (or even incorrect maintenance) which may generate
disproportionate consequences with respect to the triggering damage over the structural
lifetime.
Several hazards that could affect a structure are listed in Table 2-4, partially derived from
Starossek & Haberland (2012). It should be noted that in relation to the term ‘errors’ in
Figure 2-6, this should be interpreted that robustness serves as a last resource against
adverse events resulting from human errors.

Figure 2-6. Classification of threats

Source: Bontempi et al., 2007 (redrafted)

Table 2-4. Possible hazardous events

Impact, explosion, fire, excessive


Faults External Man-made
loading, vandalism, terrorist attack

Earthquake, landslide, extreme wind


Environmental action, tornado, heavy snow load,
scour, rock fall, volcano eruption

22
Lack of strength, cracks, ageing and
Intrinsic
deterioration

Design, material, construction and


Errors usage errors, lack or wrong
maintenance

The characteristics of exposures and hazards are very different, depending on the specific
type and time and space dependencies. Accidents, explosions and technical failures are
generally events occurring suddenly. Floods and fire storms are usually relatively slower
hazards, while deterioration processes and climate change are much slower.
The hazards to be considered in relation to structural robustness assessment can be
specified either by a relevant authority or, where not specified, on a project-specific basis
by relevant parties.10

2.2.2 Continuous damage and life-cycle robustness


The concepts of structural robustness, disproportionate failure and progressive collapse
are generally associated to damage suddenly caused by accidental and extreme actions,
such as explosions or impacts. However, damage may also arise gradually in time due to
ageing and deterioration processes (Ellingwood 2005, Biondini & Frangopol 2016). The
detrimental effects of these phenomena may lead to unsatisfactory structural performance
under service loadings. Moreover, depending on the damage propagation mechanism, such
kinds of damage could also involve disproportionate consequences. Deterioration processes
may also interact over time with damage induced by other natural or man-made hazards.
For instance, these processes may become very relevant for ageing bridges exposed to
high levels of traffic loadings and seismic actions.
During the past decades, significant attention has been devoted, in different countries, to
the condition rating of existing structures and infrastructures (Biondini & Frangopol 2018).
The economic impact of ageing and deterioration processes on these systems is
enormously high, particularly for bridges. A proper modelling of the structural system over
its entire life-cycle by taking into account the effects of deterioration processes, time
variant loadings, maintenance and repair interventions is therefore fundamental. In this
context, robustness needs to be ensured over the entire service life.
Deterioration processes caused by environmental aggressiveness are generally very
complex. Their behaviour over time depends on the damage mechanisms, the type of
materials and the structures considered. For instance, considering traditional civil
engineering materials, the lifetime of steel elements may be affected by corrosion and
fatigue. For reinforced concrete structures, chemical processes associated with
carbonation, sulfate and chloride attacks, reinforcing steel corrosion, alkali-silica reactions,
freeze/thaw thermal cycles process and mechanical processes such as cracking, abrasion,
erosion and fatigue may seriously affect the life-cycle performance of the structures
(Biondini & Frangopol 2019). Moreover, several damage mechanisms might be
simultaneously active, but progressing at different rates, including the sequence of
occurrence.
The evolution over time of the deterioration processes needs to be described by suitable
time-variant models. However, a mathematical description of the physical mechanism
underlying the deterioration process is often not available. In such cases, empirical models
can be successfully adopted (Ellingwood 2005).

10
Formulation as included in FprEN1990:2022

23
2.2.3 Damage modelling
An analytical description of damage processes may be seen as too complex to be adopted
or not always suitable to be incorporated in a robustness evaluation. However, effective
models can often be established for practical applications by directly assuming the
structural damage as a progressive deterioration of the mechanical and geometrical
properties at material, member and system levels. To this purpose, the amount of
deterioration can be defined by means of damage indices 𝛿 ∈ [0; 1] associated with
prescribed patterns of deterioration, with 𝛿 = 0 for the undamaged state and 𝛿 = 1 the
complete damaged state (Frangopol & Curley 1987, Biondini et al. 2008).
The mathematical expression of the time-variant damage index 𝛿 depends on the damage
mechanism considered and the level of complexity of the analysis. Several mechanisms,
including uniform corrosion in steel structures or crushing, cracking, abrasion and erosion
in concrete structures, can be effectively represented at the member level by a progressive
reduction of the effective resistant area of the member cross-section.
As an example, for steel members with hollow circular cross-sections having internal and
external radius 𝑟𝑖 and 𝑟𝑒 , respectively, and damage along the external layer of uniform
thickness Δ𝑟, the amount of damage can be specified by means of the following damage
index:
Δr (2-1)
δ=
𝑟𝑒 − 𝑟𝑖
In this way, proper correlation laws may be introduced to define the variation of the
geometrical properties of the cross-section, such as area 𝐴 = 𝐴(𝛿) and inertia moment 𝐼 =
𝐼(𝛿), as a function of the damage index 𝛿 .
Different patterns are needed when localized damage occurs. As an example, for corrosion
of bars in reinforced concrete structures, denoting p the corrosion penetration depth the
damage index 𝛿 can be defined as:
𝑝 (2-2)
𝛿=
𝐷0

where 𝐷0 is the diameter of the undamaged steel bar section. In turn, the percentage loss
𝛿𝑠 = 𝛿𝑠 (𝛿) = 1 − 𝐴𝑠 /𝐴𝑠0 of steel resistant area 𝐴𝑠 for a corroded bar depends on the corrosion
mechanism. In carbonated concrete with limited chloride content, corrosion tends to
develop uniformly on the steel bars along an external layer of thickness Δ𝑟, with 𝑝 = 2Δ𝑟
and 𝛿𝑠 = 𝛿(2 − 𝛿). On the other hand, in presence of significant concentration of chlorides,
corrosion tends to localize (pitting corrosion), and the relationship 𝛿𝑠 = 𝛿𝑠 (𝛿) depends on
the shape of the pit (Stewart 2009).
A damage index 𝛿 = 𝛿(𝑥) provides a comprehensive description of the spatial distribution
of damage over the structure. However, due to its local nature, it is not useful for global
evaluations of structural robustness. A synthetic global measure of damage  can be
derived at the member level or over all members of the system level by a weighted average
over the structural volume 𝑉 as follows (Biondini 2004):

∫𝑉 𝑤(𝒙)𝛿(𝒙) 𝑑𝑉 (2-3)
Δ=
∫𝑉 𝑤(𝒙) 𝑑𝑉
where 𝛿 = 𝛿(𝒙) is a damage index at point 𝒙 and 𝑤 = 𝑤(𝒙) is a suitable weight function.
Arithmetic average with constant weights functions 𝑤(𝑥) = 𝑤𝑚 = 𝑤0 can be adopted if there
are no portions of material volume playing a specific role in the damage process. Contrary,
different weights may be appropriate for different component materials in non-
homogeneous systems, such as reinforced concrete structures (Biondini 2009).
This approach can be adopted for modelling and quantifying sudden damage or continuous
damage by assuming suitable time-variant formulations of 𝛿 = 𝛿(𝑡) and 𝛥 = 𝛥(𝑡).

24
2.2.4 Damage propagation
After failure of one-member, other members may fail leading to a sequence of local failures
that propagate throughout the overall system until its collapse is reached. The mechanism
of damage propagation is generally related to the causes of the damage itself and also
depends on the system configuration. To explain the concept, two alternative propagation
mechanisms are considered here, defined as directionality-based and adjacency-based
mechanisms (Biondini & Restelli 2008). Other concepts are given in Starossek (2017).
In the directionality-based mechanism, damage propagates along the direction normal to
the axis of the first failed member. For example, with reference to the frame system shown
in Figure 2-7.a, the damage of member 1 is followed in sequence by the damage of
members 2, 3 and 4. The directionality-based mechanism is typical of damage induced by
severe loadings, such as explosions or impacts, which generally tend to propagate along
the direction of loading.
Figure 2-7. Damage propagation mechanisms: (a) Directionality-based; (b) Adjacency-based.
(a) (b)

Source: Biondini & Restelli, 2008

In the adjacency-based mechanism, damage propagates towards the members directly


connected with other members already damaged. For example, with reference to the frame
system shown in Figure 2-7.b, the damage of member 1 can be followed by the damage
of members 5 and 13. The adjacency-based mechanism is typical of damage induced by
aggressive agents, like chlorides, which generally tends to propagate through the structure
based on diffusion processes.
Considering the local definition of damage and a certain propagation mechanism, an
effective and complete damage scenario can be obtained by adopting a damage sensitive
fault-tree analysis (Biondini & Restelli 2008). With this approach, all the possible damage
paths, based on the topology of the structure and the propagation mechanism considered,
can be represented by branched networks where the level of activation of each nodal
connection is properly tuned to account the prescribed amount of local structural damage.

2.2.5 Modelling of hazards


To model natural and man-made hazards, in order to include them into a design process
and build a structure able to sustain these kinds of action, a set of parameters, functions
and models which describe the event are needed. In this process, there are several sources
of aleatory and epistemic uncertainties and all of them need to be properly considered in
a probabilistic framework for a realistic modelling.
Malicious attacks to structures, including terrorism, vandalism, or other intentional human
actions, are highly unpredictable and unknown. Important indicators of the consequences
associated to the occurrence of these events are the number of potential victims and the
role in society of the structural facility. It is easy to understand and reasonable to think
that a “symbolic” touristic building will have a larger level of exposure to terrorist attacks
than a residential building. However, due to the uncertainties involved in this type of
hazard, it is generally more efficient to adopt appropriate measures to protect the structure

25
in case they actually do occur (for instance using safety barriers) and to limit the possible
damage propagation over the structure.
Although EN1990 stipulates that structures should be designed, executed and maintained
adequately by qualified and experienced people and controlled by appropriate quality
management, the occurrence of errors and negligence during the design, execution and
operation phases (which could lead to significant and disastrous consequences) can never
be fully excluded and for such situations robustness provisions can limit the consequences
resulting from such errors or negligence.
Potential hazards may be classified according to uncertainties (Vrouwenvelder 2010):
— Known and dealt with: associated risks are accepted with no additional measures or
reduced to an acceptable level (may include natural hazards and ordinary loads);
— Known in principle, but unrecognized or ignored: generic design requirements for these
actions (as human errors in design, construction and use) are generally provided;
— Unknown or unforeseeable: no specific information is available.
The latter category can be better specified as unforeseen or unforeseeable hazards at the
time of design/assessment. The flutter mechanism of the Tacoma Narrow Bridge, for
instance, may be considered as unforeseeable at during its design. Past failure events
should serve as important lessons to be considered for future design, execution and
operation phases.

2.3 Consequences of failure


The consideration of failure consequences is a fundamental task in the evaluation and
assessment of robustness of structural systems. A robustness evaluation has to compare
the amount of damage and the corresponding structural consequences.

2.3.1 Direct and indirect consequences


Consequences are typically divided into two categories. According to the distance in space
and time from the triggering event, direct and indirect consequences may arise (Janssens
et al. 2010).
— Direct consequences: consequences associated to damage or failures of the
constituents of a system. These are all the possible immediate consequences (not
considering loss of system functionality) arising from direct effects of the hazards to
the structural system. Depending on how the structure reacts and adapts to its new
configuration after damage (e.g. trying to avoid progressive collapse with alternative
load paths), indirect consequences may occur.
— Indirect consequences: consequences related to loss of functionality of the system
caused by direct consequences. These outcomes are usually indicated as follow-up
consequences with respect to initial direct consequences. Indirect consequences can
also include environmental, psychological, political and reputational aspects and all the
managerial and delay’s costs.
In the context of consequences of failure, the concepts of structural robustness and
vulnerability play a fundamental role. Structural robustness is the ability of a system to
avoid a performance reduction disproportionately larger with respect to the corresponding
damage.
In a risk-based context, structural robustness and vulnerability can be associated with
collapse and damage probabilities as shown in Figure 2-8. In this Figure, P[E] denotes the
probability of occurrence of an accidental event E, that affects the structure; P[D|E]
represents the conditional probability of the initial/direct damage D due to the event E;
P[C|D] is the conditional probability of a structural collapse C, due to damage D, and P[C]
denotes the probability of disproportionate collapse as result of a hazard event.

26
Figure 2-8. Structural robustness and vulnerability

Source: Starossek & Haberland 2012 (redrafted)

Direct consequences may occur under hazardous events and the structural vulnerability
plays the role of avoiding initial consequences on the system.
Moreover, direct and indirect consequences also depend on the system definition
considered in the analysis. For instance, a bridge can be considered as the main system
composed by different structural members that play the role of constituents of the system.
In this case, after a hazardous event, several direct and indirect consequences may occur.
On the other hand, the same bridge can be seen as a single component of an infrastructural
network and the system will be associated to other direct and indirect consequences.
Therefore, consequences from structural failure are also function on the level of detail (i.e.
the system definition) in the risk assessment. An example related to a roadway network is
presented in Figure 2-9.

Figure 2-9. System definition for a roadway network

Source: Faber et al. 2007

27
It is worth noting that a different system definition is usually adopted for bridges relative
to other types of structures (e.g. buildings), since the former are usually part of a system-
of-systems.

2.3.2 Consequences in the Eurocodes


Structures are generally categorised in consequence classes, considering the structural
type, occupancy and size. In “Eurocode 0 – Basis of Structural Design”, three consequence
classes are introduced based on the level of economic, social and environmental
consequences. The different classes are identified to suggest appropriate design strategies
in order to increase the damage-tolerance of the structural system.
The consequence classes are associated to different accidental design situations, as stated
in “Eurocode 1 – Part 1-7: Accidental actions”, see the section 3.1.2.2.3 of this report.
The consequence classes proposed by Eurocode are assumed to be dependent on the
building type, number of storeys and the floor area. Moreover, Eurocode 1 recommends
different strategies to provide an acceptable level of robustness to sustain localised failure
without a disproportionate level of collapse, as outlined in 3.1.2.2.3.
Direct and indirect failure consequences depend on the representation of the analysed
system and its boundaries. However, consequences may be classified not only based on
their distance in time and space with respect to the triggering event. Generally,
consequences associated to building and bridge failures are divided into four categories:
human, economic, environmental and social consequences (Canisius 2011). Table 2-5
presents a list of possible direct and indirect failure consequences.
Obviously, the definitions of the considered system and its boundaries play an important
role in the categorisation of consequences enabling a clear and rational distinction between
direct and indirect consequences. For instance, management costs may become direct
consequences as result of repairs on damaged portions of a bridge.
An important issue is the comparison of different types of consequences, particularly in the
unit of measures used to quantify failure consequences (Faber & Stewart 2003). A
monetary value may be associated for example to repair, rescue and clean-up costs. Also
consequences belonging to the economical field (e.g. energy and management costs) can
be easily quantified through the amount of money used and therefore accounted into a
cost-benefit analysis. However, a monetary unit of measure may not be always a suitable
choice, in particular when environmental, loss of functionality and reputation consequences
and human loss are present. Expressing human injuries and fatalities in terms of monetary
values may be problematic for ethical reasons.

28
Table 2-5. Classification of failure consequences

Type Consequences

Human Injuries
Fatalities
Psychological damage

Economic Repair of initial damage


Replacement/repair of contents
Rescue costs
Clean up costs
Collateral damage to surroundings
Loss of functionality
Traffic delay/management costs

Environmental CO2 Emissions


Energy use
Pollutant releases

Social Loss of reputation


Loss of public confidence

Consequences are also classified according to the item affected:


— Tangible consequences: injuries, fatalities, damage or failure to structural components,
physical and functionality losses;
— Intangibles consequences: loss of opportunities and reputation, psychological effects,
deferred production, losses in environmental attributes (e.g. pollution).
Consequences of failure can also be classified according to the probability of occurrence P:
— Systemic (P=1): construction costs, decommissioning costs, etc.;
— Occasional (P<1): consequences related to frequent accidents;
— Rare (P<<1): consequences of rare accidents.

2.3.3 Factors affecting the consequences of failure


Consequences of failure depend on many factors related to the hazards, properties and
function of the structure, time-frame considered and surrounding environment
(Chryssanthopoulos et al. 2001, Sørensen 2010). A list of the dominant factors affecting
the consequences of failure is given below:
— Nature of the hazard
It is intuitive how the nature, magnitude and the duration of a certain hazard will affect
its consequences: the larger the magnitude and duration the larger will be the
associated consequences. Moreover, the type/nature of the hazard may lead to
additional consequences and risks. A gas explosion will affect the mechanical properties
of the system and moreover the nature of the hazard can also generate fumes and
toxic pollutants increasing the environmental consequences;
— Properties of the structure

29
The building typology, its age, size and layout, the choice of materials adopted, the
type and quality of construction will influence the consequences of failures. Moreover,
vulnerability and robustness are strictly correlated to the properties of the structure;
— Use of the structure:
The average number of people daily present in a building and therefore, the amount of
people exposed to the hazard, will affect the consequences of failures (such as injuries
and fatalities);
— Location of the structure:
The position of a building exposed to an accidental hazard will influence the
consequences of failures for different reasons, such as:
● Pollutants agents may have larger consequences in urban areas than buildings
in rural places, potentially increasing the consequences for human health;
● The availability of emergency services may be better in urban areas than in the
rural ones. However, the access in the latter areas is likely to be easier and less
critical than in crowded places;
● Regarding the cost of repair or reconstruction, remote locations may have higher
costs due to increased labour and material costs;
● For bridges, the type of road served by the bridge influences the traffic intensity
and therefore the amount of people possibly exposed to a hazard and the
associated traffic delay costs.
— Environmental and meteorological conditions:
During and after the exposure hazardous event, environmental and meteorological
conditions may influence the consequences of the event itself. The most intuitive
example is the quality and condition of air (for instance the wind speed and direction),
both during and after the hazardous event, may increase or decrease the environmental
consequences (leading, for example, to the dispersion of pollutant agents generated by
the hazard event);
— Actual time of the hazardous event:
General structures have different occupancy levels during the day. Work places
generally have high occupancy levels during working hours. On the other hand,
residential buildings reach high number of people at night. The same behaviour is
evident for bridges which have a certain peak of level of usage during the day. These
time considerations inevitably lead to periods during the day which the potential human
consequences (injuries and fatalities) are larger. Additional daily, weekly, monthly,
seasonally variations may influence consequences of failure for certain hazards;
— Time-frame considered:
It is important to specify the time-period adopted for the evaluation. Generally, the
time-frame should be chosen based on the consequences’ duration. When all the
consequences of a certain hazard are completed, the analysis can stop. However, for
several hazardous events and its consequences (particularly for the intangible ones) it
is difficult to identify a clear duration. Moreover, for bridge failures since their
consequences will affect the entire transportation network, the requested period to
reconstruct the bridge and offset all the consequences may be very large (years). For
these reasons it could be necessary to fix a certain time-frame for the consequence
analysis where its value is strictly related to the consequences to be considered.

30
References
Agarwal J., England, J., ‘Recent developments in robustness and relation with risk’,
Proceedings of the Institution of Civil Engineers, 161:4, 183-188, 2008.
Agarwal, J., Blockley, D.I., ‘Structural integrity: hazard, vulnerability and risk’,
International Journal of Materials and Structural Integrity, Vol. 1, pp. 117-127, 2007.
Agarwal, J., Blockley, D.I., Woodman, N.J., ‘Vulnerability of structural systems, Structural
Safety’, Vol. 25, pp. 263-286, 2003.
André J., ‘Structural robustness: A revisit’, Structural Engineering and Mechanics,
Structural Engineering and Mechanics, Vol. 76, No 2, 2020, pp. 193-205.
André, J., Beale, R., Baptista, A., ‘New indices of structural robustness and structural
fragility’, Structural Engineering and Mechanics, Vol. 56, No 6, 2015, pp. 1063–1093.
André, J., Faber, M., ‘Proposal of guidelines for the evolution of robustness framework in
the future generation of Eurocodes’, Structural Engineering International, Vol. 29, No 3,
2019, pp. 433–442.
ASCE, American Society for Civil Engineering, ‘Minimum Design Loads for Buildings and
Other Structures’, ASCE/SEI 7-16, 2016.
Baker, J.W., Schubert, M., Faber, M.H., ‘On the assessment of robustness’, Journal of
Structural Safety, vol. 30, pp. 253-267, 2008.
Biondini F., ‘A Three-dimensional Finite Beam Element for Multiscale Damage Measure and
Seismic Analysis of Concrete Structures’, 13th World Conference on Earthquake
Engineering, Paper No. 2963, Vancouver, B.C., Canada, August 1-6, 2004.
Biondini, F., ‘A Measure of Lifetime Structural Robustness’, ASCE/SEI Structures Congress
2009, Austin, TX, USA, April 30-May 2, 2009. In: Proceedings of the Structures Congress
2009, L. Griffis, T. Helwig, M. Waggoner, M. Hoit (Eds.), American Society of Civil
Engineers, USA, 2009.
Biondini, F., Frangopol, D.M., ‘Time-variant Robustness of Ageing Structures’, Chapter
6, Maintenance and Safety of Ageing Infrastructure, Y. Tsompanakis & D.M. Frangopol
(Eds.), Structures and Infrastructures Book Series, CRC Press, Taylor & Francis Group, 10,
163-200, 2014.
Biondini, F., Frangopol, D.M., ‘Life-cycle performance of deteriorating structural systems
under uncertainty: Review’, Journal of Structural Engineering, ASCE, 142(9), F4016001,
1-17, 2016.
Biondini, F., Frangopol, D.M., ‘Time-variant Redundancy and Failure Times of Deteriorating
Concrete Structures considering Multiple Limit States’, Structure and Infrastructure
Engineering, Taylor & Francis, 13(1), 94-106, 2017.
Biondini, F., Frangopol, D.M., ‘Life-cycle performance of civil structure and infrastructure
systems: Survey’, Journal of Structural Engineering, ASCE, 144(1), 06017008, 1-7, 2018.
Biondini, F., Frangopol, D.M., (Eds.), 'Life-Cycle Design, Assessment and Maintenance of
Structures and Infrastructure Systems', American Society of Civil Engineers (ASCE),
Reston, VA, USA, 2019.
Biondini, F., Restelli, S., ‘Damage Propagation and Structural Robustness’, First
International Symposium on Life-Cycle Structural Engineering (IALCCE'08), Varenna, Italy,
June 10-14, 2008. In Life-Cycle Civil Engineering, F. Biondini and D.M. Frangopol (Eds.),
CRC Press, Taylor and Francis Group, 131-136.
Biondini, F., Frangopol, D.M., Restelli, S., ‘On Structural Robustness, Redundancy and
Static Indeterminacy’, ASCE Structures Congress 2008, Vancouver, B.C., Canada, 2008.

31
Bontempi, F., Giuliani, L., Gkoumas, K., ‘Handling the exceptions: dependability of systems
and structural robustness’, Proc., 3rd International Conference on Structural Engineering,
Mechanics and Computation (SEMC), Cape Town, South Africa, pp. 104-110, 2007.
Callaway, D.S., Newman, M.E.J., Strogatz, S.H., Watts, D.J., ‘Network Robustness and
Fragility: Percolation on Random Graphs’. Physical Review Letters; 85:5468-5471, 2000.
Canisius, T.D.G., ‘Structural Robustness Design for Practising Engineers’, COST Action
TU0601 – Robustness of structures, 2011.
Canisius, T. D. G., Sorensen, J. D., Baker, J. W., ‘Robustness of structural systems – a new
focus for the Joint Committee on Structural Safety (JCSS)’, Proc., 10th Int. Conf. an
Applications of Statistics and Probability in Civil Engineering (ICASP10), Taylor and Francis,
London, 2007.
Carper, K.L. & Smilowitz, R., (Eds.). ‘Mitigating the potential for progressive
disproportionate structural collapse’, Special Issue of the Journal of Performance of
Constructed Facilities, ASCE, 20(4), 116 pp., 2006.
Clifton, J.R., and Knab, L.I., ‘Service life of concrete’, NUREG/CR-5466, 1989 (U.S. Nuclear
Regulatory Commission: Washington, D.C.), 1989.
Corley, W. G., Mlakar, P. F., Sozen, M. A., Thornton, C. H., ‘Oklahoma City bombing:
Summary and recommendations for multihazard mitigation’, Journal of Performance of
Constructed Facilities. Vol. 12, no. 3, pp. 100-112, 1998.
Chryssanthopoulos, M., Janssens, V., Imam, B., ‘Modelling of failure consequences for
robustness evaluation’. In: IABSE – IASS Symposium, London, UK, 2001.
Ellingwood, B.R., ‘Design and Construction Error Effects on Structural Reliability’, Journal
of Structural Engineering, Vol. 113 No. 2, pp. 409-422, 1987.
Ellingwood, B.R., ‘Risk-informed condition assessment of civil infrastructure: state of
practice and research issues’, Structure and Infrastructure Engineering, 1(1), 7–18, 2005.
Ellingwood, B.R., ‘Mitigating risk from abnormal loads and progressive collapse’, Journal of
Performance of Constructed Facilities, ASCE, 20(4), 315-323, 2006.
Ellingwood, B.R., Dusenberry, D.O., ‘Building Design for Abnormal Loads and Progressive
Collapse’, Computer-Aided Civil and Infrastructural Engineering, 20 194-205, 2005.
EN 1991-1-7, ‘Eurocode 1 – Actions on structures, Part 1-7: General actions – accidental
actions’, Comité European de Normalization (CEN), 2006.
Eurocode 0 - Basis of structural design, European Committee for Standardisation.
Eurocode 1 - Action on Structures, Part 1-7: General Actions – accidental actions, European
Committee for Standardisation.
Faber, M.H., Maes, M.A., Baker, J.W., Vrouwenvelder, T., Takada, T., ‘Principles of risk
assessment of engineered systems’. In 10th International Conference on Application of
Statistic and Probability in Civil Engineering (ICASP10), 2007.
Faber, M.H., Stewart, M. G., ‘Risk assessment for civil engineering facilities: critical
overview and discussion’. Rel. Eng & Sys. Safety, Elsevier, 80, 173-184, 2003.
fib, fib Bulletin 65: Model Code for Concrete Structure 2010 – Final draft, Volume 2.
Frangopol, D.M., Curley, J.P., ‘Effects of Damage and Redundancy on Structural Reliability’,
ASCE Journal of Structural Engineering, 113(7), 1533-1549, 1987.
Frangopol, D.M. & Klisinski, M., ‘Weight-strength-redundancy interaction in optimum
design of three-dimensional brittle-ductile trusses’, Computers and Structures, Pergamon
Press, 31(5), 775-787, 1989.
Frangopol, D.M. & Nakib, R., ‘Redundancy in highway bridges’, Engineering Journal,
American Institute of Steel Construction, 28(1), 45-50, 1991.

32
Ghosn, M., Moses, F., Frangopol, D.M., ‘Redundancy and robustness of highway bridge
superstructures and substructures’, Structure and Infrastructure Engineering, Taylor &
Francis, 6(1–2), 257–278, 2010.
GSA, ‘Progressive collapse analysis and design guidelines for new federal office buildings
and major modernization projects’, General Services Administration (GSA), 2003.
Gulvanessian, H., Calgaro, J. and Holický, M. (2012), Designer’s Guide to Eurocode: Basis
of Structural Design ,(2nd Edition), Thomas Telford Ltd, London, UK.
Haberland, M., Starossek, U., ‘Progressive collapse nomenclature’, Proceedings, ASCE SEI
2009 Structures Congress, Austin, Texas, April 29-May 2, 2009, pp. 1886-1895, 2009.
Haberland, M., Starossek, U., ‘Terminology for treating disproportionate
collapse’, Proceedings, The Fifth International Conference on Bridge Maintenance, Safety,
and Management (IABMAS2010), Philadelphia, USA, July 11-15, 2010.
Huber, P.J., ‘Robust Statistical Procedures’, 2nd ed. CBMS-NSF Regional Conference Series
in Applied Mathematics; 68. Society for Industrial and Applied Mathematics: Philadelphia,
67 p, 1996.
Kouvelis, P., Yu, G., ‘Robust Discrete Optimization and Its Applications. Non-convex Optimization
and Its Applications’, V14. Kluwer Academic Publishers: Dordrecht, Boston, 356 p, 1997.
Janssens, V., O’Dwyer, D. W., Chryssanthopoulos, M., ‘Building failure consequences’,
COST Action TU0601 – Robustness of structures, 2010.
Lu, Z., Yu Y., Woodman, N.J., Blockley, D.I., ‘A theory of structural vulnerability’, The
Structural Engineer, 77(18), 17–24, 1999.
Meyer, B., ‘Object-Oriented Software Construction’, 2nd edn. Prentice Hall PTR, 1997.
National Transportation Safety Board, ‘Collapse of a Suspended Span of Interstate Route
95 Highway Bridge Over the Mianus River’, Greenwich, Connecticut, June 28, 1983.
Highway Accident Report NTSB/HAR-84/03. Washington, D.C., 1984.
Okasha, N.M. & Frangopol, D.M., ‘Time-variant redundancy of structural systems’,
Structure and Infrastructure Engineering, 6, 279-301, 2010.
Pearson, C., Delatte, N. J., ‘Ronan Point Apartment Tower Collapse and Its Effect on
Building Codes’, J. Perf. of Constr. Fac., 19(2), 172-177, 2005.
Sørensen, J. D., ‘Theoretical framework on structural robustness’, COST Action TU0601 –
Robustness of structures, 2010.
Starossek, U., ‘Collapse resistance and robustness of bridges’, 4th International Conference
on Bridge Maintenance, Safety and Management (IABMAS’08), Seoul, Korea, July 13-17,
2008. In: Bridge Maintenance, Safety, Management, Health Monitoring and Informatics,
H-M. Koh and D.M. Frangopol (Eds.), CRC Press, Taylor & Francis Group, 2008.
Starossek, U., ‘Progressive Collapse of Structures’, 2nd Edition, ICE Publishing, 2017.
Starossek, U., Haberland, M., ‘Measures of structural robustness – Requirements &
applications’, ASCE Structures Congress 2008, Vancouver, B.C., Canada, 2008.
Starossek, U., Haberland, M., ‘Disproportionate collapse: terminology and procedures’, ASCE,
Journal of Performance of Constructed Facilities, Vol. 24, No. 6, pp. 519-528, 2010.
Starossek, U., Haberland, M., ‘Approaches to measure of structural robustness’, Structure
and Infrastructure Engineering, 625-631, 2011.
Starossek, U. and Haberland, M., ‘Robustness of structures’, Int. J. Lifecycle Performance
Engineering, Vol. 1, No. 1, pp.3-21, 2012.
Stewart, M.G., ‘Mechanical behaviour of pitting corrosion of flexural and shear
reinforcement and its effect on structural reliability of corroding RC beams. Structural
Safety, 31, 19 – 30, 2009.

33
Val, D.V., Val, E.G., ‘Robustness of frame structures’, Struct. Engrg. Int., 16(2), 2006.
Vrouwenvelder, T., ‘Treatment of risk and reliability in the Eurocodes’, Structures &
Buildings, 161(SB4), 209-214, 2008.
Vrouwenvelder, A.C.W.M., ‘Probabilistic modelling of exposure conditions’, Proceedings of
the Joint Workshop of COST Actions TU0601 and E55, eds. J. Köhler, H. Narasimhan & M.
H. Faber, ETHZ: Zurich, pp.45-54, 2010.
Zhu, B., Frangopol, D.M., ‘Reliability, redundancy and risk as performance indicators of
structural system during their life-cycle’. Engineering Structures, 41, 34-49, 2012.

34
3 Overview of current standardization and design guidelines

3.1 Existing national standards, regulations and practices

3.1.1 Development and evolution of current design guidelines


After the collapse at Ronan Point (1968) the first code provisions for structural robustness
were introduced in the United Kingdom with the implementation of the Fifth Amendment
of the Building Regulations (Minister of Housing and Local Government, 1970). The basic
idea of these first code provisions was that minimum levels of structural redundancy must
be ensured to provide a minimal level of robustness by member survival. This concept of
notional member removal evolved with time into different methods found in contemporary
international codes. Current requirements in international design codes are hence an
evolution of the guidelines introduced in the Fifth Amendment. Also following the collapse
of Ronan Point, research at the National Bureau of Standard (United States) was initiated
and a number of technical workshops on progressive collapse were held in the 1970s
(Stevens et al., 2011). During this period, engineers expressed their concerns regarding
the continued optimisation in structural design and the trend toward speeding execution
which may lead to reduced robustness and continuity in the structural system and hence
exposing structures to a greater risk of progressive collapse when unexpected loads occur.
In 1975, the National Building Code of Canada (National Research Council of Canada, 1975)
explicitly implemented provisions to prevent progressive collapse. As pointed out by
Ellingwood et al. (2007) the uniqueness of the Canadian code is that it provides specific
values on acceptable levels of the likelihood of an extreme event (10−4 per year or more)
for which a structure should be designed. Later between 1975 and 1995 little progress was
made in this research field.
However, failures such as the collapse of the Alfred P. Murrah Federal Building in Oklahoma
City and the collapse of the World Trade Centre Towers renewed the interest in progressive
collapse design and a number of codes and guidelines (GSA, 2003, 2013) implemented
provisions to minimise the likelihood of building collapse. After the bombing of the Alfred
P. Murrah Federal Building in downtown Oklahoma City on April 19, 1995, an Interagency
Security Committee (ISC) was established which was responsible for developing long-term
construction standards for non-military facilities that require blast resistance and other
specialised security needs.
After several studies and investigations of different progressive collapse design
approaches, the ‘alternative path’ method was implemented by the General Services
Administration (GSA, 2000). This approach requires that additional robustness is provided
in a structural frame through the addition of reserve capacity and ductility designed to
sustain gravity loads after the loss of a critical load-bearing element. Later in June 1996,
the Khobar Towers complex in Saudi Arabia, housing U.S. personnel from the Department
of Defense (DoD), was bombed. As a result, the U.S. Congress directed the DoD to develop
antiterrorism standards for the construction of military facilities. These standards were
developed to reduce the vulnerability of structures on military installations to terrorist
attacks and to improve the safety of occupants. In 2001, DoD issued interim design
guidance specifically addressing progressive collapse to clarify interim antiterrorism
requirements (DoD, 2001). This guidance adopted an alternative load path method to
reduce the risk of progressive collapse that was similar to the GSA criteria. DoD updated
its antiterrorism standards for buildings in the 2002 publication of Unified Facilities Criteria
(UFC) 4-010-01: Minimum Antiterrorism Standards for Buildings (DoD, 2002) that included
the requirement to consider progressive collapse. In 2005, DoD developed UFC 4-023-03:
Design of Buildings to Resist Progressive Collapse (DoD, 2005) to provide more specific,
enforceable guidelines to support compliance with the UFC 4-010-01 requirement. Later
the UFC 4-023-03 guideline was updated in 2009 (DoD, 2009).
In other parts of the world, explicit considerations on structural robustness were not
included in codes until the beginning of the 21 st century. In China, after the triggering
event of the 2008 Wenchuan earthquake which resulted in a large number of collapsed

35
buildings, the Code for Anti-Collapse Design of Building Structures was released in 2014
and approved by the Ministry of Housing and Urban-Rural Development of China (China
Association for Engineering Construction Standardization (CECS), 2014; Li et al., 2014).
This code contains design methods for steel and concrete buildings which align well with
other international codes. In Australia, starting in 2016 the current code has introduced
general and brief requirements based on the notional member removal and key element
design for all building classes (Australian Building Codes Board (ABCB), 2016).
In 2004, the European standard EN 1991-1-7 (CEN, 2006) was completed and received a
positive vote by the member states. The code describes the principles and application rules
for the assessment of accidental actions on buildings and bridges. The leading design
principle of this code is that local damage is acceptable, provided that it will not endanger
the structure and that the overall load bearing capacity is maintained during an appropriate
length of time to allow necessary emergency measures to be taken (Gulvanessian and
Vrouwenvelder, 2006). An overview of well-known progressive collapse events and
consecutive developments of design guidelines or standards in time is illustrated in Figure
3-1.

Figure 3-1. Timeline of the main progressive collapse events and the developments of design
provisions.

Source: Qian et al., 2016

The close relationship between design code updates and the lessons learnt from damaging
events is indicative of the dynamic character of design codes. Design codes are established
to be progressively updated in order to enclose new scientific knowledge, state-of-art best
practices in design, or lessons learned through empirical evidence. In the particular case
of the European standard, the European Commission recently issued a requirement to
amend and update the current version of the Eurocode (EC, 2012), being one of the aims
of the update the “strengthening of the requirements for robustness” and the “extension
of existing rules for robustness”. It is in-line with this requirement that the following
Section discusses the existing standardisation, practice, and theory of the robustness
provisions.

36
3.1.2 Existing European provisions for robustness
The following Section provides a comprehensive overview of the state-of-art in the
European provisions for robustness up until 2019. Provisions and guidelines included in
pre-standards are not included in this overview.

3.1.2.1 Eurocode 0 (2002) – General statements


As a basic requirement for structural robustness, Eurocode EN 1990 states the following
the clause in section 2.1:
(4)P - A structure shall be designed and executed in such a way that it will not be damaged
by events such as :
— explosion,
— impact, and
— the consequences of human errors,
to an extent disproportionate to the original cause.
NOTE 1 The events to be taken into account are those agreed for an individual project with
the client and the relevant authority.
NOTE 2 Further information is given in EN 1991-1-7.
(5)P - Potential damage shall be avoided or limited by appropriate choice of one or more
of the following:
— avoiding, eliminating or reducing the hazards to which the structure can be subjected;
— selecting a structural form which has low sensitivity to the hazards considered;
— selecting a structural form and design that can survive adequately the accidental
removal of an individual member or a limited part of the structure, or the occurrence
of acceptable localised damage;
— avoiding as far as possible structural systems that can collapse without warning;
— tying the structural members together.
(6) The basic requirements should be met:
— by the choice of suitable materials,
— by appropriate design and detailing, and
— by specifying control procedures for design, production, execution, and use relevant to
the particular project.
(7) The provisions of [..] should be interpreted on the basis that due skill and care
appropriate to the circumstances is exercised in the design, based on such knowledge and
good practice as is generally available at the time that the design of the structure is carried
out.
With the basic requirement in clause (4)P the Eurocodes indicate some exposures to
consider during the structural robustness assessment and implicitly allow that some local
damage/failure may be accepted. Moreover the emphasis is laid on a design and execution
which avoid disproportionate consequences and is based on due skill, knowledge and good
practice. Clause (5)P provides different strategies which can be applied in the design
against disproportionate collapse.
The use of the term disproportionate in clause (4)P makes the design concept clear.
However, the term is subjected to the individual interpretation of the designer and does
not allow quantifications of what is an acceptable robustness level. However, as indicated
by the text of EN 1990, there is a need to interact with the client or relevant authority of
a project to discuss the hazardous events which need be considered. Hence the hazardous

37
events which can be considered are not only limited to explosions, impact or human errors.
Note that the term ‘robustness’ as such is not explicitly defined in EN 1990.

3.1.2.2 Eurocode 1, part 1-7 (2006)

3.1.2.2.1 Scope
In addition to Eurocode EN 1990, EN 1991-1-7 gives the principles and application rules
for the assessment of identifiable and unidentifiable accidental actions on buildings and
bridges and gives a definition of robustness. Further, as a main design principle, it is stated
that local damage is acceptable as long as the structural stability and load-bearing capacity
of the building is not endangered for an appropriate period of time to allow necessary
emergency procedures to take place, e.g. the safe evacuation and rescue of personnel
from the building and its surroundings. Longer periods of survival may be required for
buildings used for handling hazardous materials, provision of essential services, or for
national security reasons.

3.1.2.2.2 Strategies
Compared to EN 1990, EN 1991-1-7 adds exposures from unidentifiable causes to the list
specified in EN 1990. In this regard, strategies applicable for identified accidental actions
and for unidentified accidental actions are outlined (Figure 3-2). In EN 1990, an accidental
action is defined as:
‘An action, usually of short duration but of significant magnitude that is unlikely to occur
on a given structure during the design working life.” A note to the definition states: “An
accidental action can be expected in many cases to cause severe consequences unless
appropriate measures are taken.’
Typical examples of identified accidental actions are fire, explosions, earthquakes, impact,
floods, avalanches, landslides and so on. Note that in the Eurocode system, fire and
earthquake are dealt with in specific parts of the Eurocodes, supplemented by country-
dependent specifications is national annexes, and treated by an alternative limit state
formulation. Regarding impact and explosion, some guidelines are given in EN 1991-1-7.
Furthermore, for concrete structures in accidental design conditions, EN 1992-1-1 provides
some additional guidance.
Once an accidental action is defined, this identified action can be dealt with by classical
safety formats (i.e. considering appropriate partial factors and load combinations) and
advanced structural analysis. However, it is important to indicate that, for the verification
of accidental design situations, no alternative target reliability levels are specified. Next to
the strategy of designing the structure to sustain an identified action, EN 1991-1-7 also
gives the possibility to prevent or reduce the action and to design the structure to have
sufficient minimum robustness. To implement the latter strategy, the following methods
can be adopted:
1. Designing certain components of the structure upon which the stability of the structure
depends as key elements to increase the likelihood of the structure’s survival following
an accidental event;
2. Designing structural members, and selecting materials, to have sufficient ductility
capable of absorbing significant deformation energy without rupture;
3. Incorporating sufficient redundancy in the structure to facilitate the transfer of actions
to alternative load paths following an accidental event.
A note is given with a reference to informative annexes A and C of EN 1991-1-7 and to EN
1992-1999 to find guidance on what is considered as ‘sufficient ductility’. However, limited
information regarding this topic is available in the so-referred Eurocodes.

38
Figure 3-2. Distinction strategies for identified and unidentified accidental actions in EN 1991-1-7

Source: CEN, 2006.

3.1.2.2.3 Annex A: Design for consequences of localised failure in buildings


Related to the basic principles for damage mitigation mentioned in Eurocode EN 1990, for
unidentified accidental actions, more general robustness requirements (for example
prescriptive tying forces) are introduced. As is the case for general design, the objective
of these requirements is to reduce the risks to an economically acceptable cost. Risk may
be expressed in terms of the probability and the consequences of undesired events. Hence,
risk-reducing measures consist of probability reducing measures and consequence
reducing measures (Gulvanessian and Vrouwenvelder, 2006). However due to the
occurrence of extreme and unexpected events, no design can be made risk free. As a
result, localised failure is acceptable to a certain extent as long as the global structural
stability is not endangered. To limit the extent of localized failure due to unidentified
accidental actions, three strategies are given in EN 1991-1-7 (CEN, 2006) (right hand side
of Figure 3-2):
 Designing the structure so that in the event of a localised failure (e.g. failure of a
single member) the stability of the whole structure or of a significant part of it would
not be endangered;
 Designing key elements, on which the stability of the structure depends, to sustain
the effects of a notional accidental action Ad;
 Applying prescriptive design/detailing rules that provide acceptable robustness for
the structure (e.g. three-dimensional tying for additional integrity, or a minimum
level of ductility of structural members subject to impact).
The design for accidental situations is of particular importance where a collapse may result
in consequences in terms of injury to human beings, or may have significant economic,
social or environmental consequences. A convenient approach to decide which structures
should be designed for accidental actions, is to classify the structures or structural
members in categories according to the consequences of an accident. As such, EN 1991-
1-7 (CEN , 2006) makes a distinction between the strategies to be applied for unidentified
accidental design situations, on the basis of the Consequence Classes defined in EN 1990
(CEN, 2015). EN 1990 classifies structures in three categories based on the consequences
of failure, see Table 3-1.

39
Table 3-1. Consequence Classes according to EN 1990.

Consequence Examples of buildings and civil


Description
Class (CC) engineering works

CC3 High consequence for loss of


Grandstands, public buildings
human life, or economic, social or
where consequences of failure are
environmental consequences
high (e.g. a concert hall)
very great

CC2 Medium consequence for loss of Residential and office buildings,


human life, economic, social or public buildings where
environmental consequences consequences of failure are
considerable medium (e.g. an office building)

CC1 Low consequence for loss of Agricultural buildings where


human life, and economic, social people do not normally enter
or environmental consequences (e.g. storage buildings),
small or negligible greenhouses
Source: CEN, 2015

In annex A of EN 1991-1-7 (CEN, 2006), the table defining the Consequence Classes (CCs)
in EN 1990 (CEN, 2015) is further extended. Furthermore, CC2 is subdivided into two
subclasses: CC2a (buildings up to 4 storeys) and CC2b (buildings up to 15 storeys).
However, this table is not exhaustive and can be adjusted by national annexes.
Subsequently, in EN 1991-1-7, the strategy to be adopted for accidental design situations
is based on the consequence classes and can be summarized as follows:
 CC1: Provided a building has been designed and erected in accordance with the
rules given in EN 1990 to EN 1999 for satisfying stability in normal use, no further
specific consideration is necessary with regard to accidental actions from
unidentified causes.
 CC2a: In addition to the recommended strategies for Consequences Class 1, the
provision of effective horizontal ties, or effective anchorage of suspended floors to
walls should be provided;
 CC2b: In addition to the recommended strategies for Consequences Class 1, the
provision of:
o horizontal ties should be provided together with vertical ties in all supporting
columns and walls.
or alternatively,
o the building should be checked to ensure that upon the notional removal of
each supporting column and each beam supporting a column, or any nominal
section of load bearing wall, the building remains stable and that any local
damage does not exceed a certain limit. Where the notional removal of such
columns and sections of walls would result in a damage extent in excess of
the agreed limit, or other such limit specified, then such elements should be
designed as a ‘key element’;
 CC3: A systematic risk assessment of the building should be undertaken taking into
account both foreseeable and unforeseeable hazards. An examination of the specific
case should be carried out to determine the level of reliability and the depth of the
structural analyses required. This may require the use of refined methods such as
dynamic analyses and non-linear models. Guidance on the preparation of the risk
analysis is given in Annex B of EN 1991-1-7 (CEN, 2006).

40
3.1.2.2.4 Prescriptive tie rules
Regarding the design of the proposed ties in the strategies presented above, two types of
ties are indicated in EN 1991-1-7 (CEN, 2006): horizontal ties and vertical ties. Next a
distinction is made between framed structures and load bearing wall structures to
determine the design tie force and placing of the ties. In Table 3-2, the different design tie
forces given by EN 1991-1-7 are summarised. It is important to note that horizontal floor
tying offers a potential resistance mechanism via tensile catenary/membrane action in the
event of loss of a column/vertical load bearing member. In this respect, it should be
emphasised that the currently prescribed tying forces should be improved for typically
expected ductility levels, and that the current rules do not incorporate any ductility
considerations. Further information about the different ties specified in EN 1991-1-7 is
presented below.

Table 3-2. Design tie forces according to EN 1991-1-7 (CEN, 2006).

Design value of tensile force


Considered
tie
Framed structures Load bearing wall structures (CC2b)

𝑇𝑝 = 0.4(𝑔𝑘 + 𝜓 𝑞𝑘 ) ∙ 𝑠 ∙ 𝐿
Peripheral ties 𝑇𝑝 = 𝐹𝑡
≥ 75𝑘𝑁

𝑇𝑖 = 0.8(𝑔𝑘 + 𝜓 𝑞𝑘 ) ∙ 𝑠 ∙ 𝐿 𝐹𝑡
Internal ties 𝑇𝑖 = max (𝐹𝑡 (𝑔𝑘 + 𝜓𝑞𝑘 ) 𝑧 )
≥ 75𝑘𝑁
7.5 5

The vertical tie should be capable


of carrying an accidental design 34𝐴 𝐻 2
( )
tensile force equal to the largest 8000 𝑡
Vertical ties 𝑇 = max
design vertical permanent and 𝑘𝑁
100
variable load reaction applied to 𝑚 𝑜𝑓 𝑤𝑎𝑙𝑙 )
(
the column from any one storey.

s is the spacing of the ties;


L is the span length of the tie;
𝜓 is the relevant factor in the expression for combination of action effects for the accidental design situation
(i.e. 𝜓1 or 𝜓2 in accordance with expression (6.11b) of EN 1990);
𝐹𝑡 is 60 kN/m or 20 + 4𝑛𝑠 kN/m, whichever is less;
𝑛𝑠 is the number of storeys;
z is the minimum of 5 times the clear storey height H or the greatest distance in metres in the direction of the
tie, between the centres of the columns or other vertical load bearing members;
A is the cross-section area in mm2 of the wall measured on plan;
H is the clear storey height of the wall;
t is the thickness of the wall.

Horizontal ties
Framed structures (CC2a and CC2b buildings)
 Horizontal ties should be provided around the perimeter of each floor and roof level
and internally in two right angle directions to tie the column and wall elements
securely to the structure of the building. The ties should be continuous and be
arranged as closely as practicable to the edges of floors and lines of columns and
walls. At least 30% of the ties should be located within the close vicinity of the grid
lines of the columns and the walls.
 Horizontal ties may comprise rolled steel sections, steel bar reinforcement in
concrete slabs, or steel mesh reinforcement and profiled steel sheeting in composite

41
steel/concrete floors (if directly connected to the steel beams with shear
connectors). The ties may consist of a combination of the above types.
 Members used for sustaining actions other than accidental actions may be utilized
for the above ties.
 The design strengths of the perimeter and internal ties are given in Table 3-2.
Load bearing wall structures
1. For CC2a buildings:
Appropriate robustness should be provided by adopting a cellular form of
construction designed to facilitate interaction of all components including an
appropriate means of anchoring the floor to the walls.
2. For CC2b buildings:
Continuous horizontal ties should be provided in the floors. These should be internal
ties distributed throughout the floors in both orthogonal directions and peripheral
ties extending around the perimeter of the floor slabs within a 1.2 m width of the
slab. The design strength of the ties can be found in Table 3-2.
Vertical ties
Each column and wall should be tied continuously from the foundation level to the roof
level. The design strength for the ties in case of framed and load bearing wall structures
can be found in Table 3-2.

3.1.2.2.5 Notional removal of load bearing elements


This strategy is proposed for CC2b buildings. It consists in checking that upon the notional
removal of each load bearing element (one at a time in each storey of the building), the
building remains stable and that any local damage does not exceed a certain limit. The
application of this method results in checking the ability of a structure to activate
alternative load paths in case of loss of a load bearing element. Limited information is
provided in EN 1991-1-7 (CEN, 2006) on how to apply this strategy.

3.1.2.2.6 Admissible damage and key element design


In case the second strategy is chosen for a CC2b building, i.e. notional removal of load
bearing elements, the stability and damage of the structure should be assessed. Where
the notional removal of a load bearing element would result in an extent of damage in
excess of the recommended limit of admissible local failure, then such an element should
be designed as a ‘key element’. This key element is supposed to be designed to withstand
a recommended accidental load Ad of 34 kN/m2 which should be applied in horizontal and
vertical directions (in one direction at a time). This recommended value was based on the
gas explosion at the collapse of Ronan Point for which the pressure was estimated to be
between 14 kPa and 83 kPa (Ellingwood et al., 2007). The recommended limit of admissible
local failure in case the notional load bearing element removal is applied, is 15% of the
floor area or 100 m2, whichever is smaller, in each of two adjacent storeys (Figure 3-3).
Unfortunately the standard does not provide detailed guidance on how to quantify the
extent of the damage. Note that the values above are recommended values and the correct
values to be used in a member state can be found in their respective National Annex.
Further note that in Figure 3.3 from EN1991-1-7, the term ‘collapsed area’ instead of
‘damage’ could be deemed more appropriate.

42
Figure 3-3. Recommended limit of admissible damage according to EN 1991-1-7.

Source: CEN, 2006

3.1.2.2.7 Annex B: Information on risk assessment


For buildings in consequence class 3, EN 1991-1-7 recommends a formal quantitative risk
analysis (CEN, 2006). To perform and execute this risk analysis, guidance can be found in
Annex B. The recommended steps for this assessment are:
1. Definition of scope and limitations;
2. Qualitative risk analysis (inventory and description);
3. Quantitative risk analysis (modelling and calculations);
4. Risk evaluation and mitigation measures;
5. Risk communication.
As indicated by Gulvanessian and Vrouwenvelder (2006), the depth and complexity of the
risk assessment should be dictated by the problem at hand. Risk analysis in a rigorous
form including extensive statistical analyses is used only in special cases. In many cases,
a qualitative analysis of risks and envisaged countermeasures should be sufficient.
In case of a formal quantitative risk analysis according to annex B of EN 1991-1-7
(CEN, 2006), following formula is given to evaluate the total risk:

𝑁𝐻 𝑁𝐷 𝑁𝑆
(3-1)
𝑅 = ∑ 𝑃[𝐻𝑖 ] ∑ ∑ 𝑃[𝐷𝑗 |𝐻𝑖 ] ∙ 𝑃[𝑆𝑘 |𝐷𝑗 ] ∙ 𝐶(𝑆𝑘 )
𝑖=1 𝑗 𝑘=1

In this equation, it is assumed that the structure is subjected to 𝑁𝐻 different hazards, that
the hazards may damage the structure in 𝑁𝐷 different ways (can be dependent on the
considered hazards) and that the performance of the damaged structure can be discretised
into 𝑁𝑆 adverse states 𝑆𝑘 with corresponding consequences 𝐶(𝑆𝑘 ). 𝑃[𝐻𝑖 ] is the probability of
occurrence (within a reference time interval) of the ith hazard, 𝑃[𝐷𝑗 |𝐻𝑖 ] is the conditional
probability of the jth damage state of the structure given the ith hazard and 𝑃[𝑆𝑘 |𝐷𝑗 ] is the
conditional probability of the kth adverse overall structural performance 𝑆𝑘 given the ith
damage state. Based on this formula, three analysis steps can be distinguished (Figure
3-4):
1. Identification and modelling of relevant accidental hazards. Assessment of the
probability of occurrence of different hazards with different intensities;
2. Assessment of damage states to the structure from different hazards. Assessment
of the probability of different damage states and corresponding consequences for
given hazards;

43
3. Assessment of the performance of the damaged structure. Assessment of the
probability of inadequate performance(s) of the damaged structure together with
the corresponding consequence(s).

Figure 3-4. Illustration of the steps in a risk analysis for structures subject to accidental actions
according to EN 1991-1-7; (a) hazard, (b) damage, (c) collapse.

Step 1 Step 2 Step 3


Identification and modeling of Assessment of damage states to Assessment of the performance of
relevant accidental hazards structure from different hazards the damaged structure

Assessment of the probability of Assessment of the probability of Assessment of the probability of


occurrence of different hazards different states of damage and inadequate performance(s) of the
with different intensities corresponding consequences for damaged structure together with
given hazards the corresponding consequence(s)
Source: adapted from CEN, 2006

Next the evaluated risk should be compared to some risk acceptance criteria which are
usually left to the member states. In annex B, limited guidance is also given for these
criteria. Basically the ALARP is mentioned, which stands for ‘as low as reasonably
practicable’. In other words, apart from a lower bound for the individual and socially
accepted risk levels, an economical optimization is recommended. Based on Equation
(3-1), the following strategies are identified to mitigate the risk:
 Reducing the probability of the hazard occurrence (i.e. reducing 𝑃[𝐻𝑖 ]);
 Reducing the probability of significant damage given the hazard (i.e. reducing
𝑃[𝐷𝑗 |𝐻𝑖 ]). This is related to the vulnerability of the structure;
 Reducing the probability of adverse structural performance (i.e. reducing 𝑃[𝑆𝑘 |𝐷𝑗 ]).
This can be obtained by providing sufficient redundancy and consequently is related
to the robustness of the structure.

3.1.2.3 Material specific rules


Material specific rules are only reflected in EN1992-1-1 for concrete structures. Such
specific rules are not reported in the other material oriented Eurocodes.

3.1.2.3.1 EN1992-1-1
For reinforced concrete structures, the Eurocodes give some additional guidelines to design
for robustness in Eurocode EN 1992-1-1 (CEN, 2005). In section 9.10 of this Eurocode
part, the following is mentioned:

44
‘Structures which are not designed to withstand accidental actions shall have a suitable
tying system, to prevent progressive collapse by providing alternative load paths after local
damage.’
Note that in this clause, no distinction is made between identified and unidentified
accidental actions as is done in EN 1991-1-7 (CEN, 2006). Moreover, the design approach
of EN 1991-1-7 (CEN, 2006) based on the categorisation of a structures into Consequence
Classes is not mentioned in this section. In section 9.10 of EN 1992-1-1 (CEN, 2005),
simple design rules are given to design a tying system incorporating the following ties:
 Peripheral ties at each floor and roof level within 1.2 m from the edge;
 Internal ties at each floor and roof level in two orthogonal directions;
 Horizontal ties for edge columns and walls;
 Vertical ties where required, particularly in panel buildings of 5 storeys or more,
continuous from the lowest to the highest level of the structure.
Further it is mentioned that the ties are intended as a minimum reinforcement and not as
an additional reinforcement to that required by regular structural analysis. The design
values for the respective ties according to EN 1992-1-1 (CEN, 2005) are given in Table
3-3. It should be noted that structural detailing rules, such as minimum reinforcement ratio
and minimum anchorage length can also contribute to achieving robustness.

Table 3-3. Design values of ties according to EN 1992-1-1 (CEN, 2005).

Considered tie Design value of tensile force

Peripheral tie(1) 𝐹𝑡𝑖𝑒,𝑝𝑒𝑟 = 𝑙 ∙ 𝑞1 ≤ 𝑞2

Internal tie(2) 𝐹𝑡𝑖𝑒,𝑖𝑛𝑡 = 20 𝑘𝑁/𝑚 (floors with screeds)


or
𝑙1 +𝑙2
𝐹𝑡𝑖𝑒 = 𝑞3 ≤ 𝑞4 (floors without screeds)
2

Horizontal ties to 𝑓𝑡𝑖𝑒,𝑓𝑎𝑐 = 20 𝑘𝑁/𝑚 (façade walls)


columns or walls
𝐹𝑡𝑖𝑒,𝑐𝑜𝑙 = 150 𝑘𝑁

Vertical tie(3) The vertical ties should be capable of carrying the load in the
accidental design situation, acting on the floor above the
column/wall accidentally lost.
(1) 𝑙 is the length of the end span in meters; the recommended value for 𝑞1 is 10 kN/m and for 𝑞2 is 70 kN;
(2) 𝑙1 and 𝑙2 are span lengths in meters on either side of the beam line; the recommended value for 𝑞3 is 20 kN/m
and for 𝑞4 is 70 kN;
(3) Other solutions e.g. based on the diaphragm action of remaining wall elements and/or on membrane action
in floors, may be used if equilibrium and sufficient deformation capacity can be verified.
Source: CEN, 2005

Note that the values of Table 3-3 are recommend values; the actual values to be used in
a member state can be found in their respective National Annex.

3.1.2.4 National Standards


In general, requirements on robustness in national standards are country-specific and it is
difficult to provide a general overview. Due to the Ronan Point accident, the British
standardisation has the longest tradition in development of the requirements on robustness
(COST, 2011).

45
3.1.2.4.1 British Standards
Since the Ronan Point collapse, the British Standards have taken the lead in stating explicit
design provisions against progressive collapse. The British Standards started with the
Building (Fifth Amendment) Regulations in 1970 (Minister of Housing and Local
Government, 1970) which were applied to all buildings having five or more storeys
(including basement storeys). The initial guidelines emphasised general tying of various
structural elements of a building together, to provide structural integrity, continuity and
redundancy. With the application of ties and ensuring continuity between structural
elements the resistance of wall panels subjected to pressure in the event of an explosion
is enhanced and the ability to bridge over a lost element is also improved. In addition,
load-bearing elements which are vital for the general building stability should be designed
as key elements, able to withstand an accidental load, i.e. a pressure of 34 kPa.
Since then, the guidelines in the UK have been evolved to the Approved Document A of
the UK Building Regulations, which is generally adopted in the Eurocodes (Eurocode 1, BS
EN 1991-1-7:2006). The design requirements are described in Annex A. While Annex A is
informative, the UK National Annex effectively makes the annex normative, stating that
the ‘guidance ... should be used in the absence of specific requirements in BS EN 1992-1-
1 to BS EN 1996-1-1 and BS EN 1999-1-1 and their National Annexes.’

3.1.2.4.2 Czech Standards


In the Czech standards, directives are provided for houses constructed of panels. For these
constructions, it is required to verify the overall spatial stiffness. The design of the tie
reinforcement in the horizontal and vertical joints prescribes a value of 15 kN/m over the
width or length of a panel house. Furthermore, reinforcement is required in each joint of
vertical and horizontal members by additional or latent ties.
For masonry structures, reinforcing bars at each floor level are required. For multi-storey
buildings, recommended construction rules are provided and the total height of the building
is restricted (COST, 2011).

3.1.2.4.3 Danish code DS 409


Robustness is introduced in the Danish Code of Practice for Safety of Structures as a
general requirement to all structures in order to reduce the sensitivity of the structure with
respect to unintentional loads and defects. The Danish code provisions are based on a
probabilistic approach and require a step-by-step procedure for all structures where
consequences of failure are serious (COST, 2011). Similar to the Eurocodes, the
requirements for robustness of structures are related to the consequences of failure of the
structure. Only for structures in the highest consequence class CC3, robustness shall be
documented (Sorensen-J.D., 2008):
— by demonstrating that those parts of the structure essential for the safety only have
little sensitivity with respect to unintentional loads and defects, or;
— by demonstrating a load case with ‘removal of a limited part of the structure’ or;
— by demonstrating sufficient safety of key elements.

3.1.2.4.1 Italian code and guidelines


The Italian building code (MIT, 2018) is fully consistent with Eurocodes’ provisions,
recommending that constructions should have an adequate level of structural robustness.
This property is therein defined as the ability of the structure to avoid disproportionate
damage with respect to the magnitude of hazardous events such as explosions and impact.
Depending on the expected occupancy of the construction and consequences of collapse,
the Italian building code expects that an adequate level of structural robustness may be
achieved through single design strategies or a combination of them. In this respect, a
proper conceptual design of the structure, including the selection of a structural shape and
typology with low sensitivity to exceptional loads or local damage, is suggested. The

46
possible use of passive or active control systems is recommended as additional measure
of risk mitigation.
The Italian code establishes that exceptional loads are those produced by events such as
fire, explosion and impact. To achieve an adequate level of robustness, the designer should
consider either the hazardous scenarios and exceptional actions that mostly influence the
design of the structure or scenarios prescribed by the client, including scenarios associated
with local damage and structural deterioration. According to European provisions,
structural robustness is assessed under an exceptional load combination. In the case of
explosions and impact, the Italian code does not provide analysis methods for assessing
the propagation of local damage throughout the structure. It is also noted that the Italian
code allows the designer to assume partial safety factors equal to unity when defining
design values of material properties for robustness verifications. This provision is amended
only in the case of masonry structures, where the partial safety factor of masonry may be
set equal to one-half of those provided for gravity load conditions, but this is likely to be
seen together with the higher partial factor adopted in the Italian code, resulting in a partial
factor higher than 1.
In case of timber structures, the Italian code states that robustness requirements may be
met through proper design choices and construction detailing, which should take into
account at least the following criteria:
— protection of the structure and its components against humidity;
— use of ductile connections;
— use of composite members with ductile global behaviour;
— limitation of timber portions subjected to tensile stresses perpendicular to fibres.
The Italian code provisions may be effectively used in combination with CNR-DT 214/2018
guidelines issued by the National Research Council of Italy (CNR, 2018). Those guidelines
provide detailed formulations for modelling of a variety of exceptional loads, including
actions caused by tsunamis, landslides, floods, eruptions, windstorms, detonations,
terrorist attacks, sabotage, and human errors in design and construction. A chapter of
CNR-DT 214/2018 guidelines deals with the risk of progressive collapse, whereas other
chapters provide recommendations for risk mitigation, conceptual design criteria, design
methods and detailing rules for several types of structural systems.

3.1.3 Provisions for robustness outside Europe

3.1.3.1 United States


In the United States, the Department of Defense (DoD) researched and then adapted in
their Unified Facilities Criteria (UFC) (DoD, 2009) what they considered the best available
approaches from different organizations, specifically the British Standards and Eurocodes
(CEN, 2006). UFC 4-023-03 employs a combination of direct and indirect design methods
and includes specific guidelines for reinforced concrete, structural steel, masonry, cold-
formed steel and timber constructions. In the first version of UFC 4-023-03, the specified
level of progressive collapse design requirement was based on the Level of Protection (LOP)
prescribed for the considered structure. LOP is a military concept and is connected to the
asset value of the structure, which in turn depends on the occupancy of the building, its
mission and other factors. Consequently, the consequences of a low-probability collapse
event are reflected by this concept. In the second version of UFC 4-023-03, the LOP
concept was replaced by Occupancy Categories to determine the required level of
progressive collapse design to make the document more general and applicable for civilian
consensus-based organisations (Stevens et al., 2011). Other improvements of the
document were based on further research performed after the 9/11 collapse to make the
design requirement more economical and practical.
In the following section, the main concepts of the UFC 4-023-03 guidelines are discussed.
For existing and new constructions, the level of progressive collapse design for a structure

47
is correlated to the Occupancy Category (OC) (Table 3-4. ). The progressive collapse design
requirements employ three design/analysis approaches: Tie Forces (TF), Alternative Path
(AP), and Enhanced Local Resistance (ELR). In the following, each design procedure is
summarized.
Tie Forces method (TF)
The Tie Forces (TF) procedure prescribes a tensile force strength of the floor or roof system,
to allow the transfer of load from the damaged portion of the structure to the undamaged
portion. Hence, with this approach, the building is mechanically tied together enhancing
continuity and development of alternative load paths. The tie forces can be provided by
the existing structural elements that have been designed using conventional design
methods to carry the standard loads imposed upon the structure. There are three
horizontal ties that must be provided: longitudinal, transverse and peripheral (Figure 3-5).
Vertical ties are required in columns and load bearing walls. Structural members such as
beams, girders and spandrels, are allowed only to carry the tie forces in case their
connections can be shown capable of carrying the required tie force magnitudes while
undergoing rotations of 0.20 radians (11.3 degrees).

48
Table 3-4. Occupancy Categories and Design Requirements.

Occupancy Design Requirement


Category

I No specific requirements.

II Option 1: Tie Forces (TF) for the entire structure and Enhanced Local
Resistance (ELR) for the corner and penultimate columns or walls at the
first storey.
OR
Option 2: Alternative Path (AP) for specified column and wall removal
locations.

III Alternative Path for specified column and wall removal locations and
Enhanced Local Resistance (ELR) for all perimeter first storey columns
or walls.

IV Tie Forces and Alternative Path for specified column and wall removal
locations and Enhanced Local Resistance for all perimeter first storey
columns or walls.
Source: DoD, 2009

Figure 3-5. Tie forces in a frame structure.

Source: DoD, 2009

If all of the structural elements and connections can be shown to provide the required tie
strength, then the tie force requirement has been met. If the vertical design tie strength
of any structural element or connection is less than the vertical required tie strength, the
designer must either: 1) revise the design to meet the tie force requirements or 2) use the
Alternative Path method to prove that the structure is capable of bridging over this deficient
element. The latter requirement does not avoid the partial collapse of the structure below
the column.
To design the ties, a Load and Resistance Factor Design (LRFD) approach is proposed, i.e.
the design tie strength is taken as the product of the strength reduction factor 𝜙 and the

49
nominal tie strength 𝑅𝑛 calculated in accordance with the requirements and assumptions
of applicable material specific codes:
(3-2)
𝜙𝑅𝑛 ≥ 𝑅𝑢

where 𝑅𝑢 = ∑ 𝛾𝑖 𝑄𝑖 is the required tie strength, taking into account the respective load
effects 𝑄𝑖 and load factor 𝛾𝑖 . For steel reinforcement in tension in reinforced concrete
elements, the strength reduction factor 𝜙 shall be taken as 0.75. The floor load 𝑤𝐹 in kN/m2
to calculate the tie forces is computed as:

𝑤𝐹 = 1.2𝐷 + 0.5𝐿 (3-3)

where D and L are respectively the dead and live load in kN/m 2. For situations with non-
uniform load distribution or large concentrated loads over the floor area, additional
guidance is given.
The design tie forces for the different types of ties are summarized in Table 3-5. . Further
guidance on the positioning and continuity of the ties for different structural systems can
be found in UFC 4-023-03 (DoD, 2009).

Table 3-5. Summary design tie forces according to UFC 4-023-03

Tie type Required tie strength

Longitudinal and transverse ties 𝐹𝑖 = 3𝑤𝐹 𝐿 [kN/m]

Peripheral ties Framed and two-way load bearing wall buildings:


𝐹𝑝 = 6𝑤𝐹 𝐿 + 3𝑊𝐶 [kN]
One-way load bearing wall buildings:
𝐹𝑝 = 6𝑤𝐹 𝐿 + 3𝑊𝐶 + 3𝑊𝑊 [kN]

Vertical ties The vertical tie must have a design strength in


tension equal to the largest vertical load received by
the column or wall from any one storey using the
floor load 𝑊𝐹 .
(1) The value for the length 𝐿 depends on the considered system, i.e. framed or load bearing wall buildings;
(2) 𝑊𝐶 is 1.2 times the dead load of cladding over the length 𝐿;
(3) 𝑊𝑊 is the dead load of a wall over a length equal to the clear storey height.
Source: DoD, 2009

Alternative Path method (AP)


When a structural element cannot provide the required tie strength, the designer may use
the AP method to determine if the structure can bridge over the deficient element after it
has been removed. For Occupancy Categories II Option 2, III and IV, the AP method must
be applied for the removal of specific vertical load bearing elements.
The requirement for the AP method is that the structure must be able to bridge over specific
vertical load bearing elements which are notionally removed from the structure. If a
structure cannot be shown to bridge over a removed element, the engineer must develop
suitable or similar re-designs. For alternative load path analyses, the acceptance criteria
include the following:
 Strength limits of the members are not exceeded;
 Deformation limits of the members and connections, expressed in terms of
deflections and rotations, are not exceeded;
 Spread of damaged members is limited.

50
To perform the alternative path analyses, a general LRFD philosophy is employed taking
into account a load combination for extraordinary events (Table 3-6) and resistance factors
to define design strengths. Three analysis procedures can be used, ordered according to
increasing complexity: Linear Static (LSP), Non-linear Static (NSP) and Non-linear Dynamic
(NDP). Modifications and guidance for each method and material-specific structure
acceptance criteria for the analyses are given to accommodate the particular issues
associated with progressive collapse. For instance, for the linear static procedure, load
increase factors depending on the applied material and structural system are given to
account for non-linear effects.

Table 3-6. Load combination for alternative load path analysis according to UFC 4-023-03.

Applied analysis Load combination after notional member removal

Linear static Floors above removed column:


analysis
𝐺𝐿𝐷 = Ω𝐿𝐷 [1.2𝐷 + (0.5𝐿 𝑜𝑟 0.2𝐿)]
Floors away from removed column:
𝐺 = 1.2𝐷 + (0.5𝐿 𝑜𝑟 0.2𝐿)

Non-linear static Floors above removed column:


analysis
𝐺𝑁 = Ω𝑁 [1.2𝐷 + (0.5𝐿 𝑜𝑟 0.2𝐿)]
Floors away from removed column:
𝐺 = 1.2𝐷 + (0.5𝐿 𝑜𝑟 0.2𝐿)

Non-linear dynamic Complete structure


analysis
𝐺𝑁𝐷 = 1.2𝐷 + (0.5𝐿 𝑜𝑟 0.2 𝑆)
𝛺𝐿𝐷 and 𝛺𝑁 are respectively the load increase factors for linear static and non-linear static analysis.
Source: DoD, 2009

Enhanced Local Resistance method (ELR)


The enhanced local resistance is required for some Occupation Categories and has as
objective to ensure that a ductile failure mechanism can form when the column or wall is
loaded to failure. To meet this objective and hence to reduce the probability and extent of
initial damage, the column or wall must not fail in shear prior to the development of the
maximum flexural strength. To check the latter requirement, a LRFD approach is used and
flexural and shear demands are defined. More information is available in UFC 4-023-03
(DoD, 2009).

3.1.3.2 Canada
The National Building Code of Canada requires structures to be designed for sufficient
structural integrity to withstand all effects that may reasonably be expected to occur during
the service life. Commentary C on Part 4 advises designers to consider and take measures
against severe accidents with probabilities of occurrence of approximately 10 -4/per year or
more which is distinct from most other national Standards in giving a specific quantified
threshold on the likelihood of the extreme event for which structures should be designed.
While the concept of placing a quantified threshold on the likelihood of events which are to
be considered is in itself valid, quantifying the likelihood of the initiating event if terrorism-
related is difficult at best due to the influence by external socio-political factors which
fluctuate according to governmental policy and international events (Arup, 2011).

51
3.1.3.3 Australia and New Zealand
Australian requirements are given as a functional statement with the requirement for the
capability of the building to withstand combinations of loads and other actions to which it
may reasonably be subjected. Associated performance requirements include resistance at
an acceptable level of safety to the most adverse combinations of loads that might result
in potential for progressive collapse (Australian/New Zealand Standards, 2002).
AS/NZS 1170.0 2002 Structural design actions – General principles states that all parts of
the structure shall be interconnected with ties capable of transmitting 5 percent of the
ultimate dead and imposed loads. The supplementary document AS/NZS 1170.0 Supp
1:2002 Structural design actions – General principles – Commentary states that:
‘The design should provide alternate load paths so that the damage is absorbed and
sufficient local strength to resist failure of critical members so that major collapse is
averted. ... Connections ... should be designed to be ductile and have a capacity for large
deformation and energy absorption under the effect of abnormal conditions.’
The materials design standards contain implicit consideration of resistance to local collapse
by including such provisions such as minimum strength, continuity, and ductility (Arup,
2011).
Bita et al., 2019 survey robustness provisions in both the National Building Code of Canada
(NBCC), ASCE-7, and Australian and New Zealand AS/NZS 1170.0 2002, and highlight that
a common limitation that is found in both is that design provisions for “disproportionate
collapse”, despite existing, are only partially subjective and not explicit.

3.1.3.4 Hong Kong


The Hong Kong Building Authority uses locally-developed codes of practice for the
structural use of steel and concrete. The approach to structural robustness, accidental
damage and disproportionate collapse essentially follows the principles and methods
adopted in the United Kingdom, although there is little specific reference to robustness in
the Hong Kong Building (Construction) Regulations or in Hong Kong Codes of Practice for
structural design. The code Structural Use of Steel 2005 issued by the Building Authority
gives guidance on the principle of design against disproportionate collapse, requiring
elements to be tied together horizontally and vertically, and for the building to be designed
to survive the removal of non-key elements by establishing alternative load paths. Key
elements which have a critical influence on the overall strength or stability of the structure,
should be ‘...designed to resist abnormal forces arising from extreme events.’
The Code of Practice for the Structural Use of Concrete 2004 more closely reflects BS 8110-
1:1997, presenting tying requirements consistent with UK practice. The general principle
is given that ‘...a structure should be designed and constructed so that it is inherently
robust and not unreasonably susceptible to the effects of accidents or misuse, and
disproportionate collapse.’ No guidance is given on any requirement for alternative load
paths or design of elements critical to the stability of the overall structure (Arup, 2011).

3.1.3.5 International Organization for Standardization


In the international context, a particular mention should be made to the updated ISO 2394
(ISO, 2015). This standard addresses decision-making in design and assessment of
structures. ISO2394 - Annex F encloses extensive robustness provisions based on risk-
based design that are significantly more in-line with what is identified as state-of-art
practice for robustness assessment.
In its risk-based approach the ISO 2394 uses 5 Classes that are comparatively more
comprehensive in characterisation of structures than the ones presented in the Eurocodes.
These use a more detailed description of class categories, and more important, provide a
quantitative indicator of risk based on the expected number of fatalities. For structures up
to Class 3, pre-specified robustness provisions are given in the code which involves some
of the methods already mentioned, such as the event control method or the alternative

52
path. For Class 4 and 5, risk-based assessment for robustness is provided in Annex F.4. In
addition to the reference to progressive or disproportionate collapse, it also uses the term
“damage insensitivity” to relate the need for comprehensive robustness provisions.

3.2 Assessment of current design guidelines


Worldwide building standards contain specific provisions for the design against progressive
collapse or they provide more general procedures to increase the structural integrity and
robustness. Depending on the country, the design guidelines can have a normative/legal
or informative status. They all emphasize the need for a good structural layout,
redundancy, ductility and continuity in order to design for progressive collapse and/or to
avoid disproportional damage. In general, three types of strategies can be identified in the
structural guidelines:
 Non-structural strategies such as minimising the exposure to hazards by preventive
measures;
 Indirect design methods providing strength, redundancy, continuity and ductility by
the use of prescriptive rules;
 Direct design methods which include the alternative path method (ability to bridge
over local damage zones) and the specific load resistance or key element method
(strengthening vital structural components).
The decision to apply a certain strategy depends on the associated consequences in case
of collapse whether using Consequence Classes (i.e. British Standards and Eurocodes) to
categorise the structure or by using Occupancy and Risk Categories (UFC 4-023-03). Note
that building standards that address progressive collapse, by a direct or an indirect
approach, usually contain specific requirements for tying systems as well as requirements
to check the structural stability under specific load combinations which take into account
structural damage or accidental loads. In the following Sections, the latter is compared for
the British Standards, Eurocodes and UFC 4-023-03 guidelines of the American Department
of Defence. Note that the British Standards and Eurocodes also allow to perform a complete
risk analysis in order to assess the structural robustness of a design.
EN1990 gives general guidance on non-structural measures, by recommending the
avoidance, elimination or reduction of the hazards to which the structure can be subjected
as a basic requirement for design. However, apart from the case when a full risk
assessment approach is required (e.g., EN1991-1-7 Class 3), no more detailed non-
structural measures or requirements to be used in design are given. Objectively, non-
structural strategies do not increase the resistance of a structure to disproportionate
failure. Nonetheless, more detailed measures such as,
 preventive and protective planning in site planning, or consideration of barriers, and
 restrictive planning and operation, with avoidance of non-design-specified activity
are expected to contribute to alleviate the design acceptance criteria to avoid
disproportionate collapse (Canisius et al. 2011).

3.2.1 Indirect Design methods


In all considered building standards, some prescriptive tie rules are provided to ensure the
integrity of the structure and to allow the development of alternative load paths. In most
codes, a distinction is made between horizontal ties (including perimeter ties, internal ties
and ties to columns and walls) and vertical ties. However, when comparing the different
standards, different design forces can be found for the respective ties. For example, the
value for the intensity factor for the longitudinal and transverse ties in UFC 4-023-03 is
considerably greater compared to the values for peripheral and internal ties prescribed in
EN 1991-1-7, see section 3.1.2.2.4 of this report. Even between the two Eurocode parts
EN 1992-1-1 and EN 1991-1-7, conflicting design strengths can be found. Moreover, the
British Standards and Eurocodes only indicate some design tie forces whereas the UFC
guidelines also indicate some deformation criteria which should ensure the ductility to allow

53
for the activation of the tying system. Next to the design strength of the ties, also the
placing of the ties differs between the different codes. Further, UFC 4-023-03 makes a
clearer distinction for the prescriptive tie rules between the different applied materials by
using material-dependent strength reduction factors.

3.2.1.1 Derivation of formulas for horizontal ties EN 1991-1-7


Although in all considered building standards some prescriptive tie rules are provided to
ensure the integrity of the structure, no background documentation can be found in
literature related to the derivation of these tie rules. In the following, a possible derivation
method for the prescriptive tie rules of the internal horizontal ties for framed structures
according to EN 1991-1-7 (CEN, 2006) is illustrated (Vrouwenvelder, 2008b).
In the following derivation, a framed structure is considered for which the span lengths of
the bays in both orthogonal directions are equal to 𝑠 and 𝐿 respectively (Figure 3-6). Next,
it is assumed that a central column is removed in order to activate the horizontal ties,
resulting in a central deflection 𝛿. Further it is assumed that the load initially carried by the
removed column 𝑅 is transferred by four internal ties in both orthogonal directions in
agreement with EN 1991-1-7 (CEN, 2006) which is recommending to place the internal
ties orthogonal in both directions. It is also assumed that the same tensile load is
developing in both directions.

Figure 3-6. Justification for the prescriptive tie rules for internal horizontal ties of framed
structures according to EN 1991-1-7

Neglecting the ductility and deformation limits of the internal elements, one apparently
assumes a central deflection 𝛿 of ± 16 % of the sum of both span lengths:

𝑠+𝐿 (3-4)
𝛿=
6

Next, taking into account the vertical equilibrium in this deformed state (see Figure 3-7),
the following equations can be derived:

𝑅 = 4 ∙ 𝑇𝑖 ∙ sin(𝛼) (3-5)

or

𝑅 𝛿
sin(𝛼) = ≈ (3-6)
4 ∙ 𝑇𝑖 𝑋

54
where X is the average span length of the bays in both orthogonal directions:

𝑠+𝐿 (3-7)
𝑋=
2

Figure 3-7. Vertical equilibrium for the deformed state to derive the prescriptive tie rules for
internal horizontal ties of framed structures according to EN 1991-1-7

Taking into account the accidental load combination according to EN 1990 (CEN, 2015),
the load initially carried by the removed column 𝑅 can be assumed as equal to:

𝑅 = (𝑔𝑘 + 𝜓 ∙ 𝑞𝑘 ) ∙ 𝑠 ∙ 𝐿 (3-8)

Finally, combining Equations (II.6), (II.7), (II.8) and (II.9), the internal tie forces 𝑇𝑖 can
be written as:

𝑇𝑖 = 0.78 ∙ 𝑅 ≈ 0.8 ∙ 𝑅 = 0.8 ∙ (𝑔𝑘 + 𝜓 ∙ 𝑞𝑘 ) ∙ 𝑠 ∙ 𝐿 (3-9)

which corresponds to the formula as prescribed by EN 1991-1-7 (CEN, 2006) to design the
internal horizontal ties of framed structures. Regarding the derivation discussed in previous
paragraph, following remarks should be made:
 As a simplification, the ductility and deformation limits of the elements are
neglected in this simplified approach. In case the span lengths 𝑠 and 𝐿 are both
equal to 6 m, this approach would result in a central deflection of 2 m, which is for
instance unrealistic for reinforced concrete elements;
 Considering the vertical equilibrium in a deformed state, large central deflections
results in smaller tie forces. As a consequence, the assumption of a large central
deflection could result in an unsafe design of the tie forces;
 The assumed angle of rotation is considered to be fixed and unrealistically large.

3.2.2 Direct Design methods

3.2.2.1 Alternative load path method


In case of higher associated failure consequences or in case the prescriptive tie rules from
an indirect design method cannot be fulfilled, an alternative load path method is
recommended. In each building standard, this alternative load path method is executed by
considering the notional removal of a load bearing element and assessing the residual
strength or bridging capabilities of the damaged structure. An example of the notional
removal for the alternative load path method is given in Figure 3-8, where the removal of
a column is investigated accordingly to (GSA, 2013), which is built on the UFC 4-023-03
guidelines. On one hand, in case of the Eurocodes and British Standards, little guidance is
given on how to implement this procedure. In these documents, design with the alternative
path method is not addressed as a comprehensive measure of design improvement, but
instead, it is highly emphasised on the identification of critical elements. Such characteristic
is rapidly identified when addressing the differences encountered in the codes in regard to
key element design (see for example comparative usage of key element presented in
Section 3.2.2.2). André and Faber (2019) highlighted before the strong element-centricity
of the present design provisions in the design codes. For the UFC 4-023-03 guidelines, on
the other hand, much more guidance is given such as:

55
 The position of the load bearing elements to be notionally removed;
 The load combinations to be considered to analyse the damaged structure;
 The analysis procedures and assumptions;
 The acceptance criteria for the strength and deformation capacities of the damaged
structure, depending on the used material and structural layout.
It should be mentioned that a sudden removal of the load bearing element is considered
here, thus allowing for dynamic effects.

Figure 3-8. Alternative Path method and strategy for the notional removal of elements in the
structure.

Source: GSA, 2013

In the context of the alternative path method, and to emphasize the review presented, it
is of relevance to mention the work of André and Faber (2019) that objectively highlights
requirements for the update on the alternative path method implementation in the design
codes. For example the consideration of one or more of the following provisions:
 Provision of structural redundancy (with adequate deformation and loading
capacity and resistance of elements and connections, including reserves);
 Provision of secondary load carrying mechanisms (provided that elements and
their connections are both sufficiently resistant and ductile to allow for additional
structural loading and deformations);
 Provision of structural integrity (element continuity and ductility).

3.2.2.2 Key element design


In case the damage acceptance criteria are exceeded in the British Standards and
Eurocodes, load bearing elements have to be redesigned as key elements considering a
certain accidental load. On the contrary, the key element design method or enhanced local
resistance method (Specific Load Resistance method) in the UFC 4-023-03 is applied for
higher Occupation Classes to ensure that a ductile failure mechanism can form. Further in
case the damage acceptance criteria of UFC 4-023-03 are not fulfilled, the structure should
be redesigned. As a consequence of this subtle difference, the analysis procedures of the
key element design method differ between these building standards. Little guidance is
given in the British Standards and Eurocodes on how to design a load bearing element
under the proposed accidental load. In addition, the recommended accidental load of 34
kN/m² appears to be a safe estimation of overpressure associated to a blast loading; the
use of such a high value of accidental load is not necessarily appropriate for any structures
for which key element design applies.

56
One of the aspects of key element design that is of relevance to highlight is that it may be
suitable and cost-effective for structures with a limited number of identifiable key
elements, and of interest when alternative load paths do not exist. However, its
performance is highly susceptible to the future scenarios of loads since failure in design
cannot be avoided with certainty, and this particularity of key element design should be
highlighted and considered in any future key element design provisions.
Another characteristic of the key element design in the current design codes is that it
neglects second-order effects that may be of relevance when a structure suffers a localised
failure. In particular, being a key element designed for survivability, it will not be
uncommon for, in the circumstance of failure of one key element, other neighbour
structural members to be badly damaged. Provisions of multi-damage scenarios are of
interest in robustness provisions that involve key element design. Whether or not this is
considered admissible relates to whether this damage is considered disproportionate.

3.2.3 Systematic Risk assessment


It was highlighted in Section 3.1 that EN 1991-1-7 recommends systematic risk
assessment for structures falling in CC3, with a detailed procedure being presented in
Annex B. It distinguishes a qualitative and a quantitative risk analysis.
In qualitative risk analysis all hazards and corresponding hazard scenarios should be
identified. Identification of hazards and hazard scenarios is a crucial task to a risk analysis.
It requires a detailed examination and understanding of the system. For this reason a
variety of techniques have been developed to assist the engineer in performing this part
of the analysis (e.g. PHA, HAZOP, fault tree, event tree, decision tree, causal networks,
etc.).
In the quantitative part of the risk analysis, probabilities should be estimated for all
undesired events and their subsequent consequences. The probability estimations are
usually at least partly based on judgement and may for that reason differ substantially
from actual failure frequencies. If failure can be expressed numerically the risk may be
presented as the mathematical expectation of the consequences of an undesired event.
In risk evaluation, acceptance and mitigation, should follow a procedure to decide when a
risk is identified, whether mitigating measures should be specified. In acceptance the
ALARP (As Low As Reasonable Practicable) principle is used. For mitigation different
technique are proposed, such as, hazard control or controlled collapse. A step of
reconsideration is enclosed and it involves reviewing the procedure until acceptance is
achieved. The final step is then to communicate the results and conclusions to all
stakeholders (specifying e.g., analysis, sources, assumptions or further
recommendations). EN 1991-1-7 Annex B also provides more detailed analysis of
applications to civil engineering structures.
In the context of systematic risk assessment it is of relevance to highlight two
particularities: first, this more comprehensive analysis targets a niche of structures, which
may generate discussion on the threshold that separates for example the class 2b and 3
and their disassociation with quantified risk (e.g., by comparison, classes of ISO 2394
consider risk–assessment for both top classes, one of which is more representative of the
CC2b EN 1991-1-7 Consequence Class); and second, accordingly to (Canisius, et al.,
2011), quantitative measures are those that are of interest to achieve comprehensive
frameworks (comparable ranking) for decision-making in structural design.

3.3 Flaws and inconsistencies in current standardization


Based on the discussion above, the following shortcomings of the actual building standards
or design guidelines can be identified:
 No consensus can be found for the indirect design method using prescriptive tie
rules. Neither the design tie force formulas nor the rules for the spacing of the ties
are consistent between the different codes;

57
 Guidance on the detailing of the ties is missing;
 The background of some formulas to calculate the tie forces is not always clear.
Some design tie forces of the Eurocodes can be derived from simple equilibrium
equations for a deformed state of the structure (see Section 3.2.1.1) but the
associated deformations lead to unrealistic deformation demands;
 In the Eurocodes, no specific or quantified requests in terms of deformation capacity
and ductility is reported while such properties are identified as a key issues when
considering the structural robustness;
 Current indirect design guidelines do not make a clear distinction between different
construction methods. Nonetheless different construction methods such as precast
constructions will require specific prescriptive tie rules. Specific guidance for the
design of precast structures against progressive collapse are given in the fib bulletin
63 (2012). However, this bulletin mainly adopts the recommendations according to
EN 1991-1-7. Further, an alternative method is presented and illustrated in this
bulletin to calculate the prescriptive tie forces for precast structures which assumes
unrealistic deformations;
 In the British Standards and the Eurocodes, little guidance is given on how to
perform the notional column removal or alternative load path design;
 The accidental key element design load, is based on a domestic gas explosion, but
should be applicable for other exposure scenarios as well; it should also enclose
considerations on second order effects of failure;
 Little guidance is given on how the accidental key element design load is transferred
to the key element (i.e. as a pressure to column or considering tributary area from
adjacent walls);
 With regard to the recommended 15% limit of admissible damage, the criteria on
how to decide which area is affected are unclear. Also, no background document
for the recommended value of 15% is provided and it is not clear whether this value
is applicable to all types of structures;
 The guidance to perform a complete risk assessment is insufficient which limits the
applicability of this method;
 Design formats presented in British Standards and the Eurocodes still rely to a large
extent on individual element safety classifications;
 Analysis of the CC allows to perceive that some ambiguity can be encountered in
the classification in design, in particular due to the lack of a quantified measure to
separate CC;
 Robustness should be (at minimum) an intrinsic property of the structural system.
It is a fact that considering provisions in EN1991-1-7 ensures some level of
robustness in design to fundamental actions, nonetheless, there is interest in a
more enveloping analysis on whether, and in which situations, further
considerations should be enclosed for fundamental actions;
 It seems unlikely to be possible to avoid prescriptive provisions in design codes. In
these cases, it is important to accompany robustness provisions with considerations
on the assumptions and adequacy to the structure being analysed. Design
considerations to apply when using prescriptive provisions can appear in the form
of sensitivity analyses with respect to the prescriptive variables and other design
assumptions;
 No methods are given to quantify the structural robustness of a design. Hence the
framework to obtain a uniform robustness level is missing. Since robustness
requirements aim to ensure adequate structural performances at system level, any
applicable rules should be based on the performance of structural systems. The
importance of performing a rational identification of event scenarios, damages,
element failures, collapse and consequences (direct or indirect) is lacking.

58
References
Adam, J.M., Parisi, F., Sagaseta, J., Lu, X. (2018) Research and practice on progressive
collapse and robustness of building structures in the 21st century. Engineering Structures
173, 122-149.
André, J., Faber, M., ‘Proposal of guidelines for the evolution of robustness framework in
the future generation of Eurocodes’, Structural Engineering International, Vol. 29, No 3,
2019, pp. 433–442.
Agarwal, J., England, J. (2008) Recent developments in robustness and relation with risk.
In: Proceedings of the Institution of Civil Engineers - Structures and Buildings 161(4), 183-
188.
Arup (2011) Review of international research on structural robustness and disproportionate
collapse. Report, Department for Communities and Local Government – Centre for the
Protection of National Infrastructure.
Australian/New Zealand Standards. (2002) AS/NZS 1170.0 2002 Structural design actions
– General principles.
Australian Building Codes Board (ABCB) (2016) National Construction code (NCC). Council
of Australian Governments.
Bita, H. M., Huberb, J. A., Voulpiotisc, K., Tannert, T. (2019). Survey of contemporary
practices for disproportionate collapse prevention. Engineering structures, 109578.
Canisius, T., Baker, J., Diamantidis, D., Ellingwood, B. F., Holicky, M., Markova, J.,
Vrouwenvelder, A. (2011). COST Action TU0601: Structural Robustness design for
practising engineers.
CEN (2005) EN 1992-1-1: Eurocode 2 - Design of concrete structures - Part 1-1: General
rules and rules for buildings (+AC:2010). Brussels (Belgium).
CEN (2006) EN 1991-1-7: Eurocode 1 - Actions on structures - Part 1-7: General actions -
Accidental actions. Brussels (Belgium).
CEN (2015) EN 1990: Eurocode 0 - Basis of structural design (+AC:2010). Brussels
(Belgium).
CNR (2018) CNR-DT 214/2018: Istruzioni per la valutazione della robustezza delle
costruzioni. National Research Council of Italy, Rome (Italy).
China Association for Engineering Construction Standardization (CECS) (2014) CECS 392:
Code for anti-collapse design of building structures, Beijing (China).
Department of Defense (DoD) (2001) Interim Antiterrorism/Force Protection Construction
Standards - Progressive Collapse Design Guidance. Washington, DC (US).
Department of Defense (DoD) (2002) UFC 4-010-01: Minimum antiterrorism standards for
buildings. Washington, DC (US).
Department of Defense (DoD) (2005) UFC 4-023-03: Design of buildings to resist
progressive collapse. Washington, DC (US).
Department of Defense (DoD) (2009) UFC 4-023-03: Design of buildings to resist
progressive collapse. Washington, DC (US).
Department of Defense (DoD) (2016) Review of UFC 4-023-03: Design of buildings to resist
progressive collapse. Washington, DC (US).
Droogné D. (2019) Reliability-Based Design for Robustness: Evaluation of Progressive
Collapse in Concrete Structures Taking into Account Membrane Action PhD Thesis, Ghent
University, Belgium.

59
EC. (2012). M/515 – Mandate for amending existing Eurocodes and extending the scope
of structural Eurocodes. Brussels, Belgium: European Commission (EC), p. 8. Report No.:
M/515.
Ellingwood, B.R., Smilowitz, R., Dusenberry, D.O., Duthinh, D., Lew, H.S., Carino, N.J.
(2007) Best practices for reducing the potential for progressive collapse in buildings. NIST
Internal Report 7396. U.S. Department of Commerce and National Institute of Standards
and Technology, Gaithersburg, MD (US).
fib (2012) fib Bulletin 63: Design of precast concrete structures against accidental actions,
Fédération internationale du béton (fib), 85p.
GSA (2000) Progressive collapse analysis and design guidelines for new federal office
buildings and major modernization projects. General Service Administration (GSA),
Washington, DC (US).
GSA (2003) Progressive collapse analysis and design guidelines for new federal office
buildings and major modernization projects. General Service Administration (GSA),
Washington, DC (US).
GSA (2013) Alternative path analysis and design guidelines for progressive collapse
resistance. General Service Administration (GSA), Washington, DC (US).
Gulvanessian, H., Vrouwenvelder, T. (2006) Robustness and the Eurocodes. Structural
Engineering International 16(2), 167-171.
ISO. (2015). ISO2394 (Fourth edition) - General principles on reliability for structures.
International standardization Organization.
MIT (2018) D.M. 17.01.2018: Aggiornamento delle «Norme tecniche per le costruzioni».
Ministry of Infrastructure and Transportation, Rome (Italy).
National Research Council of Canada (1975) National building code of Canada. Canadian
Commission on Building and Fire Codes.
Qian, K., Li, B., Tian, Y. (2016) Recent Progress in Understanding of Load Resisting
Mechanisms for Mitigating Progressive Collapse. ACI Special Publication 309-06, 1-18.
Sørensen, J. D. (2008), Robustness of Structures: Danish Approach. Proceedings of COST
actions ‘Robustness of structures', 159-168.
Stevens, D., Crowder, B., Sunshine, D., Marchand, K., Smilowitz, R., Williamson, E.,
Waggoner, M. (2011) Department of Defense Research and Criteria for the Design of
Buildings to Resist Progressive Collapse. Journal of Structural Engineering 137(9), 870-
880.
Vrouwenvelder, T. (2008a) Treatment of risk and reliability in the Eurocodes. In:
Proceedings of the Institution of Civil Engineers – Structures and Buildings 161(4), 209-
214.
Vrouwenvelder, T. (2008b) Eurocode 1 Accidental Actions. Downloaded on April 29, 2019
from http://eurocodes.jrc.ec.europa.eu.

60
4 State-of-the-art design considerations, approaches and
strategies for improving robustness11

4.1 General
As explained in the previous Chapters, when designing for robustness, it should be
adequately assured that no accidental and/or exceptional events or damage to the
structural members would result in disproportional consequences for the structural system,
or even total collapse of the whole structure during its lifetime. Therefore, the robustness
of the system should be adequately and appropriately considered. Design for robustness
should be an integral part of the conceptual design phase of the structure, taking into
account different strategies that can be applied. Therefore, it should be considered in the
early stages of the design process.
In general, the design of a structure for robustness can involve the – explicit or implicit –
identification of:
— Hazards H;
— Local (direct) damage D;
— Systemic damage S (follow-up/indirect damage), encompassing progressive collapse;
— Direct and indirect consequences.
In a risk-based context, the total risk Rtot in relation to accidental and/or exceptional events
can be based on the following equation (4-1), considering the addition of risk associated
to local (direct) damage and systemic (follow-up) damage, enabling to identify the
influencing factors. Note that this expression is similar to the previously discussed equation
(3-1). However, equation (4-1) is more general than what is currently proposed in
EN 1991-1-7.

𝑅𝑡𝑜𝑡 = ∑ ∑ 𝐶dir,𝑖𝑗 𝑃[𝐷𝑗 | 𝐻𝑖 ]𝑃[𝐻𝑖 ] + ∑ ∑ ∑ 𝐶ind,𝑖𝑗𝑘 𝑃[𝑆𝑘 |𝐷𝑗 ∩ 𝐻𝑖 ]𝑃[𝐷𝑗 | 𝐻𝑖 ]𝑃[𝐻𝑖 ] (4-1)
𝑖 𝑗 𝑖 𝑗 𝑘

where:
— P[Hi] the probability of occurrence of hazard Hi

— P[Dj|Hi] the probability of (direct) damage Dj conditional on hazard Hi


— P[Sk |Dj ∩ Hi] the probability of systemic damage Sk conditional on the damage Dj
and hazard Hi
— Cdir the direct consequences
— Cind the indirect consequences

Considering the expression (4-1), the following strategies for reducing the risk of
disproportionate collapse can be distinguished, which will be discussed more elaborately in
the next Section:
— Reducing one or more of the probabilities P[Hi] of the occurrence of hazards;
— Reducing one or more of the probabilities P[Dj|Hi] of (direct) damage;
— Reducing one or more of the probabilities P[Sk |Dj ∩ Hi] of systemic damage;
— Reducing direct Cdir and indirect Cind consequences.

11
Except for section 4.2, this chapter was partly drafted in coordination with fib TG3.1 developing the draft text
proposals for the fib Model Code 2020, as well as the additional information provided by WG6.PT1 and
WG6.PT2 of CEN/TC250/WG6.

61
4.2 Design strategies to prevent disproportionate collapse
First of all, it is appropriate to underline that the following strategies can be adopted in the
design and execution process to prevent or reduce the progressive collapse of the system
and/or to increase local safety. However, they do not necessary increase the robustness
of the system.
The design strategies aimed to prevent disproportionate collapse have been largely
investigated in literature (Ellingwood & Leyendecker 1978, Gross & McGuire 1983,
Dusenberry & Juneja 2002) and they can be classified based on the acceptable level of
probability of occurrence.
— Design strategies which prevent the occurrence of abnormal and dangerous events (i.e.
influencing P[Hi] in equation (4-1)). This strategy aims at reducing the exposure of the
structure to hazards. A typical approach is the event control strategy which decreases
the exposure of a structure by reducing the probability of occurrence and/or the
intensity of abnormal and hazardous events. Safety barriers against impact or base
isolators to limit the intensity of a seismic action are examples of such solutions. Event
control is a non-structural measure which does not involve structural robustness.
— Design strategies which prevent the occurrence of an initial damage (i.e. influencing
P[Dj|Hi] in equation (4-1)). This strategy aims at improving the local component
behaviour providing a sufficient local resistance in consequence of the occurrence of
abnormal events. The vulnerability of the system can also be decreased by a specific-
local-resistance method (e.g. detailing) on critical elements. In this design procedure,
aimed to prevent disproportionate collapse, the concept of structural robustness is
involved. However, the key element design approach does not necessary enhance the
robustness of the system. The possible failure of critical/key elements may promote
progressive and disproportionate collapses leading to unsatisfactory robustness levels.
— Design strategies which prevent disproportionate spreading of damage (i.e. influencing
P[Sk |Dj ∩ Hi] in equation (4-1)). This strategy is aimed at the global system behaviour
so that the spread of the initial local damage remains limited. The alternative load path
approach redistributes the forces originally carried by the failed members in the
damaged system trying to avoid a progressive failure. However, if the local damage
produces an overloading in the remaining structure, the alternative load path procedure
may promote progressive collapse and reduce the robustness of the system. The
existence of alternative load paths enhances the structural redundancy. An alternative
is the structural compartmentalization approach through which the spreading of failure
is prevented or limited by isolating the initially damaged portion of the system.
Methods to prevent progressive and disproportionate collapse can also be divided into
direct and indirect design approaches, as discussed in Section 4.5. Direct design methods,
like the local resistance approach and alternative load path method, try to verify explicitly
the collapse resistance of the structure when subjected to hazard scenarios through specific
structural analysis. On the other hand, indirect design methods try to guarantee implicitly
the collapse resistance through prescriptive design rules and recommendations.

4.3 Design considerations


A structure should be designed to have an adequate level of robustness so that, during its
design service life it will not be damaged by adverse and unforeseen events, such as the
failure or collapse of a component or part of a structure, to an extent disproportionate to
the original cause (prEN 1990:2020). To this aim, masses, stiffness and member capacities
should ideally be uniformly distributed both in plan and in elevation. Vertical members, like
columns and walls, should ideally run (without interruption or strong capacity reduction)
from the foundations to the top of the building. Also, in order to obtain the maximum
capacity in terms of ductility of diaphragm elements, similar span values should be adopted
along the longitudinal and transversal direction of the in-plan view of the building. Design
for robustness should also consider all possible actions a structure might be exposed to,

62
and the severity level of the consequences due to failure of a structural member or of a
limited part of the structure. These topics are further discussed in the Sections below.

4.3.1 Classification of accidental actions


The actions considered for designing structures for robustness are either identified
accidental actions (associated to identified accidental events) or accidental actions
associated with unidentified hazardous events. These can then be further classified as in
Table 4-1. While this is a comprehensive list, what is to be considered has to be decided
between the engineer and the client/owner, guided also by applicable regulations.

Table 4-1. Classification of identified accidental actions and accidental actions associated with
unidentified hazardous events.

Identified accidental actions Accidental actions associated with unidentified


hazardous events

Sub-class Sub-sub-class Sub-class Sub-sub-class

Impact Internal impact Actions Horizontal action, applied on limited


extent of structure (single
component)

External impact Horizontal action, applied on large


(Ground Level) extent of structure

External impact Vertical action, applied on a limited


(Above Ground extent (single component) of
Level ) structure

Explosion Internal Explosion - Vertical action, applied on large


Deflagration extent of structure

Internal Explosion – Simultaneous vertical and horizontal


Detonation action
On a limited extent (to reflect an
explosion, internal or external) or a
large extent, for example, to reflect
a ground movement

Ageing and Indirect action – global or local


deterioration extent (as a result of a deterioration
process to a larger extent than what
was accounted for at the design)

External Explosion – Defects and Indirect action – local extent (from a


Detonation Errors defect in construction or error in
design)

4.3.2 Consequence Classes


The design considerations for robustness should be based on the Consequence Class (CC)
of the structure. Consequence Classes categorise the severity level of the (direct and
indirect) consequences due to failure of a structural member or of a limited part of the
entire structure. Consequence Classes are typically defined based on the type of structure,
profile of occupancy and functionality.

63
For each design requirement, recommended for each CC, detailing rules should be coherent
with the basic assumptions of the adopted design method. Detailing rules should ensure
ductile failure modes of structural members, characterised by the formation of plastic
hinges in beams/joints or yield lines in continuous slabs, taking into account possible
deterioration in view of durability. Therefore, a ductile failure mode should be activated
prior to a brittle failure mode, such as shear failure modes in beams and slabs or punching
shear failure modes in slabs. Adequate transversal reinforcement in beams and slabs or
adequate slab thickness have to be designed to obtain this hierarchy of failure mode in
structural members.
The qualitative definition for CC in case of robustness can – when considering the difference
in severity of possible follow-up consequences – take basis in the CC categorization as
applied for reliability differentiation in EN1990, see Consequence Classes CC1 (Low), CC2
(Medium) and CC3 (High) in Chapter 3. Structures complying with the criteria for more
than one CC should be designed considering the most severe CC. The different CC are
presented in Table 4-2 alongside the definition of the Design Approaches in terms of analysis
and design methods.
According to the work of WG6.PT1 four Design Approaches are provided: DA-1, DA-2a,
DA-2b and DA-3, where the level of complexity of the Design Approaches increase from
DA-1 to DA-3. These can be linked to the CC1, CC2 and CC3 classes. The specificities of
each design approach are elaborated further in Section 4.4.
— In DA-1, no particular robustness provisions are required if the structure is designed
according to the partial factor design philosophy specified in the Eurocodes for identified
non-accidental hazardous events (i.e. persistent, transient and seismic design
situations in the EN 1990 terminology) and the consequences of a potential collapse of
the structure are acceptable taking into account the hazardous events (i.e. also
accidental and unidentified hazardous events for which the structure is not designed
explicitly – good practice design and detailing rules may provide an unknown
robustness level with respect to these two hazardous events).
— DA-2a and DA-2b apply to similar groups of structures but exhibiting different levels of
magnitude of consequences for a given design scenario. Therefore, different levels of
sophistication of the analysis and design methods should be applied to DA-2a and DA-
2b. A possible discrimination between the two design approaches can relate to:
*) DA-2a involving a design according to prescriptive rules solely (enabling the possible
activation of horizontal ties and ensuring sufficient anchorage of the suspended floors
to the wall), whereas DA-2b requires in addition also vertical tying.
*) The discrimination between the two design approaches can be as such that approach
DA-2a involves a simpler analysis and design methods than the approach DA-2b. For
example, linear elastic models may still be appropriate in DA-2a to design for
robustness with respect to all possible design situations (involving identified and
unidentified hazardous events), whereas in DA-2b non-linear analyses are deemed
necessary, together with an explicit consideration of the structural performance with
respect to relevant design situations (involving identified and unidentified hazardous
events) by structural analysis and also of the subsequent assessment of the adequacy
of the achieved robustness level.
— In DA-3, the use of risk-based methods for the verification of robustness are considered
necessary due to the significant magnitudes of consequences for a given design failure
mode scenario. Reliability-based methods may also be used, where appropriate.

64
Table 4-2. Robustness design approaches.

Robustness design approaches


ID Analysis method Design method
DA-1 No particular provisions are necessary with respect to
robustness.
DA-2a Analyses can be based on Design for robustness
simplified models of loads and using direct and/or
structural behaviour. indirect design
methods.
DA-2b Simplified models still possible.
Dynamic and/or non-linear analysis
models may become relevant.
Analyses may however be based
on simplified models of loads and
structural behaviour.
DA-3 Risk-based or reliability-based design for robustness.

In general, CCs deal with the severity of the consequences given an adverse event
occurrence and they have been used traditionally in design codes to define the target
reliability level implicit in the structural design (see EN1990). In cases where the
consequences are not negligible and the triggering adverse event is uncertain or unknown,
it may be necessary in design codes to specify supplementary design provisions, namely
design for robustness as they become more important. Moreover, particular design
considerations may be required for existing structures.

4.4 Design approaches

4.4.1 Definition of design scenarios


The existing general approach to the design of robust buildings can be either deterministic
or, as allowed for by the design equations of codes, semi-probabilistic. Commonly the risks
are considered implicitly and approximately by the use of various classifications of
structures. However, on certain occasions, the involved risks are considered explicitly when
designing for robustness, for example for buildings in the Consequence Class 3 of
Eurocodes.
Design for robustness consists of the identification and assessment of design scenarios.
Herein, each design scenario relating to a set of events or conditions occurring during the
construction or lifetime of the structure leading to a state of the system for which the effect
of disproportionate consequences should be assessed. Design scenarios can be identified
based on specified accidental actions, notional damage or loads and should be discussed
with the relevant stakeholders.
In case of design for identified hazards, the design scenarios consider prescribed actions,
such as accidental actions, and/or condition states and the resulting effect on the structure.
The accidental actions and events to be assumed in the design can differ substantially from
project to project. Therefore, it is difficult to specify the design scenarios in a standardised
manner, in particular when threat-specific design is used.
On the other hand, when possible threats cannot be specified, they are expressed by
notional actions or notional damage. Herein, notional damage is typically specified as cases
of initial local failure. Possible threats also include known threats that cannot be quantified
and are to be discussed an agreed upon between the relevant parties.

65
4.4.2 Design for specified accidental actions
When designing a structure with specific threats in mind, it is required to identify and
quantify all abnormal events that could possibly affect the structure and the resulting
actions on it. In general cases, such input data are usually incomplete and imprecise
because some threats are unforeseen, combined with the issue that their magnitude is
difficult to predict. Therefore, the application of a direct design method against (identified)
accidental actions may in some cases require to be complemented by elements of threat-
unspecific design (via direct or indirect design methods), in particular by the assumption
of notional damage. In cases where threat-specific scenarios are simplified by threat-
unspecific scenarios, attention should be paid to the definition of the design scenarios,
which should not deviate from the likely damage scenarios the structure would be
subjected to by the occurrence of identified accidental actions. For example, if the identified
accidental action is the impact of a vehicle on two columns, it is not adequate in general
to design the structure for the case where only one column is notionally removed. Note
that robustness design for identified accidental actions should preferably consider direct
design methods and threat-specific scenarios.
In general, a structure should be designed considering the worst case scenario from both
threat-specific and threat-unspecific accidental actions. The European design code
EN1991-1-7 provides guidelines for the modelling of impact loading and explosions
occurring inside buildings, which are accidental actions with high destructive potential.
Furthermore, several types of impact on buildings and bridges are considered. Reference
values and locations of impact forces are provided, that are modelled as either equivalent
static loads or dynamic loads depending on the Consequence Class of the structure. In
case of explosions, EN1991-1-7 takes into account both pressures directly acting on
structural members and pressures transferred to structural members from non-structural
members.
When structures are checked for specific accidental actions and/or to resist local damage,
load combination rules should reflect the low probability of concurrence of the accidental
action and the design live loads, i.e. partial factors which are lower than for the case of
ultimate limit state verifications.

4.4.3 Design for actions from unspecified threats

4.4.3.1 Notional damage scenarios


The analysis of the structural system under a notional damage scenario should be
performed considering the accidental load combination. Of course, when the performance
objective that no initial damage shall occur is specified, notional damage cannot be
specified as a hazard scenario and only the design methods based on event control,
protection and local resistance are applicable. Notional damage scenarios can include both
notional removal scenarios and notional deterioration scenarios.

4.4.3.1.1 Notional removal scenarios


Notional removal scenarios, consist of notionally removing structural elements and
consequently checking the structure for disproportionate consequences, e.g. applying an
alternative load path design or the identification of key elements. Preliminary structural
analyses may be used to identify the critical elements to consider to be notionally removed.
These scenarios are usually considered to be related to the failure of a connection or
structural member due to an unidentified hazard.
Structural elements to be notionally removed can include one or several columns, one or
several more panels or a nominal wall length and any other elements judged vital to the
structural performance. These members are typically, but not exclusively, perimeter
columns and/or load bearing walls between the ground and first levels. Alternatively, these
may also include interior load bearing elements in vulnerable locations.

66
The analysis of the structural system under a notional removal scenario is commonly
executed through a static analysis. In cases where dynamic behaviour is dominating the
structural response (e.g. referring to the structural response of the intact system), the
analysis should consider dynamic effects in a simplified (e.g. energy-balance based) or
extensive way.

4.4.3.1.2 Notional deterioration scenarios


In case of notional deterioration scenarios, the geometrical and/or material properties of
one or more structural elements are notionally reduced and the structure is checked for
disproportionate consequences.

4.4.3.2 Notional loads


Notional loads are generally specified as a uniformly distributed equivalent static load. A
value often referred to in standards for a notional load for buildings is 34 kN/m2. This
nominal value was recommended after the Ronan Point (UK) accident, triggered by a gas
explosion. The value should however be treated with caution in case the design scenario
considered is considerably different, e.g. in case of the design of a key element, especially
in case where the failure of the key element relates to large consequences. Moreover,
design for notional loads may be completely inefficient if the effects of unidentified
actions/influences are not covered by such notional loads.
These loads should be applied in the most unfavourable direction for the element under
consideration.

4.5 Strategies for improving robustness

4.5.1 General
Design strategies for robustness can generally be divided into direct design methods and
indirect design methods. The direct design methods explicitly aim to limit the effect of local
failure. They require structural analyses in order to evaluate the performance of the
structure for a certain damage scenario and can start from an alternative load path strategy
and/or a consequence reduction strategy. In case the previous approaches do not lead to
an adequate level of robustness considering reasonable investments, the design could also
be focused on reducing the probability of failure of critical elements by means of event
control strategy and/or the specific load resistance strategy. The indirect design methods,
on the other hand, do not explicitly consider the ability of a structure to sustain an
abnormal load effect, but aim to enhance the robustness implicitly, e.g. through the use of
prescriptive horizontal and vertical tie reinforcement.
The adoption of a particular strategy for designing a given structure for robustness may
lead to a conceptual solution with structural features which may be beneficial for some
hazard scenarios but detrimental for others, depending on the structural system, the
abnormal triggering-event, the magnitude and location of the initial failure or the type of
collapse. Therefore, in many cases, an appropriate way for economically meeting all
structural safety and robustness requirements should be based on a combination of
different design strategies as appropriate. For this purpose, a pragmatic approach can be
adopted (Tanner & Hingorani 2019):
— Adoption of continuous structural systems with a ductile behaviour;
— For hazard independent scenarios:
o Provision of either alternative load paths via prescriptive tying forces for indirect
design methods, or, notional scenarios that simulate failure of selected key
members for direct design methods;
— For hazard dependent scenarios:

67
o Provision of alternative load paths preferably making use of direct design
methods via explicit analyses, or, alternatively via prescriptive tying forces for
indirect design methods.;
— If alternative load paths are provided and verified for the case of structural key member
failure where the associated collapsed area Acol can be deemed negligible, then human
safety criteria can be neglected;
— If predefined collapse mechanisms (fuses) are inbuilt and verified, or in the case of
non-redundant structural systems (e.g. statically determinate structures) the collapsed
area can be readily established then human safety criteria should be applied for design
or assessment of key element.
— It should be noted that also horizontal stiffness measures can have a beneficial effect
on the robustness of structures, e.g. as the result of bracings, moment-resisting
frames, infill masonry walls, etc.

4.5.2 Alternative load path strategy


The alternative load path strategy explicitly considers the resistance to progressive collapse
(i.e. ‘indirect’ or ‘follow-up’ failure) when the level of damage is specified. The strategy
thus examines the situation that one or more structural elements have been damaged and
have no more load bearing capacity, and focuses on reducing the probability P[S|D∩ H].
For the remaining part of the structure, it is required that, for a specified short period of
time, the structure withstands the associated actions with an acceptable probability of
failure. An effective application of the alternative load path strategy typically consists of
providing sufficient redundancy/integrity, durability and ductility.
In frame type structures, the resisting mechanisms which can minimize the risk of
progressive collapse can include:
— Bending of the beam where the column has failed; this mechanism is generally
ineffective and seldom adopted since the beams have to be over-dimensioned;
— Vierendeel behaviour of the frame over the failed column;
— Arch effect of the beams where the column has failed; this mechanism is effective in
case the neighbouring structure is sufficiently stiff to limit horizontal displacements;
— Catenary/tensile membrane behaviour of beams/slabs, bridging the damaged column
by means of large rotations and displacements;
— Contribution of non-structural elements. For buildings, these can be as external infill
walls and partitions.
The transition of bending action in beams and slabs to tensile membrane (or catenary)
action is commonly considered to be a building’s last line of defence against progressive
collapse. If the designer relies on one of these resisting mechanisms for robustness, these
should be demonstrated through the analyses and design considerations.

4.5.3 Consequence reduction strategy


The consequence reduction strategy aims to limit unacceptable (disproportionate)
consequences (Cind and/or P[S|D∩H]) associated with damages D. The consequence
reducing measures should be selected on the basis of their risk reduction and the related
costs necessary to achieve it. These measures can include:
— Structural segmentation/compartmentalization, through which horizontal progression
of collapse can be limited effectively by dividing the structure into independent
structural systems by means of so-called ‘structural fuses’. The type and location of
structural fuses should be appropriately chosen and suitably designed, detailed,
executed and verified. Structural fuses are particularly effective in large, low buildings
but less effective for tall buildings. In case of the latter, compartmentalization generally
involves the installation of strong floors intermittently over the height of the building.

68
— Changing the context of the structure. This can be related to architectural and structural
variables such as the shape and static scheme of the building, organisational variables
such as relocation of key business operations to more protected and/or robust parts of
the building, Self-rescue and rescue by others and backup facilities.
— Other non-structural mitigation measures, which can include the organisation of an
efficient emergency response including fire-fighting teams, police, rescue teams,
nearby hospitals, exercises, feedback of experience.
Segmentation offers an alternative strategy where the spreading of failure following initial
damage is prevented or limited by isolating the failing part of a structure from the
remaining structure by so-called segment borders (Starossek, 2007; Starossek &
Haberland, 2012).
The most common form of segmentation relies on weak segment borders, allowing failure
of a specific segment without progression of failure to adjacent segments. In this mode,
segmentation acts as a fuse (Starossek & Haberland, 2012), where the extent of a segment
that is allowed to collapse would need to be determined by the design engineer in
consultation with the client and/or local authority, depending on the importance and type
of the structure.
Another form of segmentation relies on very strong segment borders which would be
designed to arrest an incipient collapse (Starossek & Haberland, 2012). In this mode,
segmentation can offer an alternate load path, typically with resistance to local damage
achieved at small deformations, or it can arrest the collapse of part of the structure.

4.5.4 Event control strategy


The event control strategy consists of preventing the occurrence of the hazard, or, limiting
the occurrence rate of the hazard to an acceptable level. This requires that the hazard (or
spectrum of hazards) is identified. The strategy focuses on reducing the probability P[H]
and does not increase the intrinsic resistance of a structure to unacceptable damage.
Measures associated with event control can include:
— Changes in the building site or access to it, for example through avoiding high risk
areas;
— Restricting the use of the structure. This might consist of using certain areas for specific
utilizations or prohibiting the storage or transport of explosives or other hazardous
material sources;
— Installation of early-warning systems for the hazard under consideration, through
active monitoring of wind, gas or fire development;
— Installation of passive measures, such as barriers to prevent vehicle collision, dykes or
fire insulation;
— Quality management to prevent human errors;
— Maintenance.

4.5.5 Key element design strategy


The key element design strategy aims at preventing or limiting the local damage (often
referred to as ‘direct damage’) caused by a certain hazard, such as blast pressures.
Therefore, the key element design strategy is a hazard specific approach focussing on
reducing the probability P[D|H].
Key elements are structural elements and/or connections which are essential to the
resistance of the structure. These include structural members on the lower floor levels that
are closest to exterior vehicle threats, piers of continuous bridges and cables in cable
supported structures. Measures associated with specific load resistance might include
increasing the element resistance and providing appropriate durability, increasing the

69
element stiffness, and using active or passive isolation techniques such as base isolation
of the structure.
Failure of a key element typically results in significant consequences since, in its absence,
the structure is usually unable to develop adequate alternative load paths, unless a
combined design strategy is applied. For example, the key element design strategy can be
supplemented with providing sufficient structural ductility. Herein structural ductility can
be achieved through providing ductility at the level of the system, the elements and the
material. Material ductility can be achieved by material strain-hardening and/or by material
deformation capacity, while ductile detailing can be achieved by using continuous bottom
reinforcement over supports, confinement at joints and adequate ties to allow for load
transfer.

4.5.6 Prescriptive rules


The application of prescriptive design rules is an indirect design method. These design rules
are not performance based and are assumed to enhance the robustness implicitly. The
most common prescriptive rules relate to providing horizontal and/or vertical tension ties.
More specifically, in framed structures, common prescriptive rules consist in providing
horizontal and/or vertical ties and for load bearing wall systems in providing effective
anchorage of floors and roofs to walls. Lastly, prescriptive rules with a more explicit nature
could be used based on analytical, numerical, empirical models.

4.6 Design for robustness against ageing and deterioration


Structural robustness should be evaluated not only with reference to accidental and
abnormal loadings leading to sudden damage, such as explosions or impacts (Ellingwood
2006), but also considering continuous damage associated with the effects of ageing and
deterioration processes. In fact, deterioration processes, such as corrosion in steel or
reinforced concrete structures, may lead over time to unsatisfactory structural
performance under service loadings and involve disproportionate effects and alternative
load redistribution paths (Biondini & Restelli 2008, Biondini 2009, Okasha & Frangopol
2010, Zhu & Frangopol 2012). Moreover, the lack of maintenance and repair activities may
also exacerbate these effects and lead over time to systems with insufficient robustness.
These effects are particularly relevant for buildings and bridges exposed to corrosion and
other kind of environmental damage. Notable events of bridge collapses due to the
environmental aggressiveness and related phenomena, such as corrosion and fatigue,
include for example the Silver Bridge in 1967 (ASCE 1968), and the Mianus River Bridge
in 1983 (NTSB 1983). It is hence necessary to consider and ensure structural robustness
over the entire system life-cycle. Quality control procedures, including visual inspections,
non-destructive tests and diagnostic activities carried out over the structural lifetime to
plan maintenance and repair interventions, may significantly reduce the probability of
initiation and propagation of damage associated with ageing and deterioration.
Design guidelines and standards on structural robustness against ageing and deterioration
are not available. The definition of damage scenarios involving deterioration processes is
a critical task depending on the system location and exposure. Ageing and deterioration
effects can be identified based on specific material damage phenomena (e.g. concrete
carbonation, steel corrosion and fatigue, among others) or notional damage scenarios (fib,
2006). Deterioration processes are generally complex phenomena but they can be
described using empirical models where geometrical and mechanical properties of the
ageing system are properly reduced in time (Ellingwood 2005). Notional damage scenario
can be also considered by means of prescribed patterns of structural deterioration applied
at cross-sectional, member, and/or system level (Biondini & Restelli 2008).
In general, the design of robust structures based on very strong key members, i.e. playing
a disproportionate role in the structural system, should be carefully considered during the
conceptual design phase, accounting for deterioration effects. Adequate solutions should
be adopted to properly protect the most important members against occurrence of

70
damage. Moreover, the degree of static indeterminacy should be adequately selected in
relation to the expected amount of damage, since an increase in the degree of static
indeterminacy does not necessarily lead to an increase of structural robustness (Biondini
et al. 2009, Biondini & Frangopol 2014).
The effects of maintenance and repairs activities in terms of structural robustness have to
be considered and planned during the design phase focusing on preventing the occurrence
and the spreading of initial damage. Structural health monitoring can also be used to
capture the occurrence and the evolution of damage processes. Time-variant robustness
measures incorporating information from structural health monitoring are hence necessary
to plan eventual repair interventions and maintenance actions to protect, improve and/or
restore the lifetime system performance.
Finally, it is worth noting that progressive decay of structural performance under ageing
and deterioration is generally affected over time by significant uncertainties related to
material properties, damage mechanisms, and decision-making processes associated with
maintenance and repair policies. A proper modelling of all these uncertainties is therefore
of essence to design and maintain robust systems when severe ageing and deterioration
effects are expected (Biondini & Frangopol 2016).

4.7 Multi-hazard design considerations


In many cases, structural design should allow multiple performance objectives to be met
against different hazards, which implies the activation of strongly different behavioural
modes and resisting mechanisms to reach target safety levels under several loading
conditions (see e.g. Section 4.4). In this respect, current Eurocodes do not define target
safety at structural system level because target reliability is associated with the most
critical local failure mode due to the most unfavourable load combination. This means that
structural safety at system level is only verified for the envelope of effects of all scenarios.
It is also noted that common design practice looks at different scenarios and local failure
modes independently, so the cumulative effect of several hazardous scenarios on the safety
level is usually ignored. Therefore, more refined design/assessment procedures could be
based upon a multi-hazard design framework, as emphasised by some papers (see e.g.:
Li et al., 2011; Adam et al., 2018) and mentioned in some building codes (see e.g. EN1991-
1-7 Annex B). In such a context, designers are thus requested to find solutions that allow
a satisfactory performance of the structure with respect to different criteria, either in case
of new constructions or when designing retrofit interventions for existing constructions.
For instance, Corley (2004) and Hayes et al. (2005) assessed the effects of alternative
seismic design and strengthening of the Murrah Federal Building on its progressive collapse
resistance, respectively. Those studies pointed out an influence of seismic resistance on
structural robustness. Hence, it is important to check at least different performances
(seismic resistance, fire resistance, robustness, etc.) independently because designing for
a specific performance can positively influence the design for another performance. Design
solutions (either explicit or implicit through prescriptive rules) to accommodate a certain
design criterion might not be suitable for another criterion (e.g. catenary/membrane
action, rotation capacity without necessarily maintaining moment resistance, etc.).
Provided that the structure is definitely designed or retrofitted to sustain gravity loads in
order to meet performance objectives under serviceability and ultimate conditions, there
may be several instances where both other loads and structural robustness must be
considered. Indeed, multi-hazard conditions may arise from the need to design, assess or
retrofit the structure against earthquake actions and identified/unidentified extreme
hazards such as fire, explosions, impact, and damage to a single element or a limited sub-
system.
Multi-hazard design could be related to
(1) either individual hazards or
(2) interacting hazards, including cascade events such as fire after blast, landslide or
tsunami after an earthquake, or vehicle impact after riverine/coastal flood. The

71
scenarios to be considered have then to be agreed for a specific project. It is
important to notice that this is currently not common practice, although the
consideration of such combined hazards might be important to consider in the
design.
In case of interacting hazards, classical single-hazard-oriented design methods may not
meet performance requirements, resulting in huge difficulties for actual implementation of
design/retrofit solutions in engineering practice. In addition, interaction with environmental
hazards involving continuous damage associated with ageing and structural deterioration
should be carefully considered over the system life-cycle since they can exacerbate
disproportionate damage effects induced by extreme events and favour damage
propagation and progressive collapse (Biondini et al. 2014). The following sub-sections
focus on two recurrent multi-hazard conditions where structural robustness must be
ensured together with earthquake or fire resistance.

4.8 Seismic design versus robustness


When the structure must be designed or retrofitted to withstand both earthquakes and
other extreme events, the engineer is requested to find a multi-hazard solution accounting
for the main differences between the effects of earthquake ground motion and those of,
for instance, heavy damage to a single structural component (e.g. a corner column of a
framed structure). The differences between seismic and robustness designs include, but
are not limited to, the following:
3. Identification and modelling of hazards and their action on the structure (impossible
to do for unforeseeable events);
4. Modelling of gravity loads (combination, dynamic amplification, etc.);
5. Roles of large deformations and floor system response;
6. Definition of performance objectives; and
7. Response of structural components and systems to earthquake ground motion and
local abnormal actions.
Seismic ground motion involves the whole base of the structure, inducing forces (especially
in the horizontal direction) and deformations throughout the structural system with some
potential concentrations of demand, particularly in case of irregularity in plan and/or
elevation. In most cases, extreme events such as impact or blast strike a limited portion
of the structure, which in turn may activate a partial or global response (especially in the
vertical direction) after that local damage occurs. This latter occurrence is responsible of a
possible dynamic amplification of gravity loads in those parts of the structure that mostly
contribute to activating the progressive collapse resistance. The response of structural
components and systems to extreme events is often characterised by large strain rates
and, more importantly, very large deformations of structural members and floor systems
that produce additional resisting mechanisms (e.g. arch and catenary action in beams,
membrane action in floors). This is not usually observed, and hence neglected, in nonlinear
response of structures subjected to seismic actions.
It should be noted that, as a matter of principle, the identification and definition of
structural systems (frame systems, wall systems, dual systems, large panel systems)
within building codes reflects a single-hazard rationale. For instance, Eurocode 8 (EN1998-
1:2004) provides a classification of structural systems with respect to their earthquake
resistance, whereas no definition is given for the case of accidental loss of vertical
components or, more in general, towards progressive collapse resistance and robustness.
One of the most interesting points of discussion in the literature is whether and how
earthquake resistance may produce suitable levels of robustness (Hayes et al., 2005;
Pekau & Cui, 2006; Gurley, 2008; Tsai & Lin, 2008; Parisi & Augenti, 2012; Livingston et
al., 2015; Lin et al., 2016). On one hand, ductility requirements are expected to improve
both seismic performance and structural robustness. By contrast, possible conflicts
between seismic and robustness designs may originate from capacity design criteria,
particularly the strong-column/weak-beam (SCWB) design rule according to the hierarchy

72
resistance criterion in current seismic design provisions. This could lead to systems that
cannot be able to provide the required performances for robustness. Nevertheless, at least
two alternative strategies can be implemented to improve both earthquake resistance and
robustness: (i) to increase the ultimate bending moment of beams and, consequently, that
of columns; and (ii) to provide weak beams with a sufficient overstrength by catenary
action, considering the key role of ultimate deformation of reinforcing steel and
reinforcement bond in beam-column joints or adjacent beams (Yu & Tan, 2014). There is
no doubt that increasing beam strength may violate the SCWB capacity design rule.
Conversely, weak columns may produce soft-storey mechanisms under horizontal actions,
resulting in a high probability of pancake collapse unless falling of upper storeys is arrested
during their impact on lower floors (see e.g. Lalkovski & Starossek, 2016). It is worth
noting that capacity design criteria are usually assumed as time-invariant in seismic design.
However, the system ductility and hierarchy of member strengths – and hence the energy-
dissipating failure mode claimed for a capacity design of the structure according to the
SCWB rule – may change over time depending on the environmental exposure of the
structure (Biondini & Frangopol 2008), possibly shifting from a typical ‘beam sway’ to a
‘column sway’ collapse mechanism (Biondini et al. 2011). This highlights the importance
of a proper combination of seismic and environmental hazards in the evaluation of the life-
cycle seismic performance and structural robustness of deteriorating systems.
In the case of steel structures, Park & Kim (2010) carried out a pushdown-based fragility
analysis to assess the progressive collapse potential of frames with different types of
connections. Xu & Ellingwood (2011) investigated the robustness of seismically designed
pre-Northridge steel moment-resisting framed buildings, which did not develop a
significant catenary action because of a high probability of connection failure.
Dealing with reinforced concrete structures, Brunesi et al. (2015) characterised the
progressive collapse fragility of European, low-rise, framed buildings designed according
to Eurocodes through incremental dynamic analysis. Both structures designed to gravity
loads only and structures designed for earthquake resistance were investigated, indicating
a significant impact of seismic design criteria and detailing on robustness. That study was
further remarked by a huge amount of pushdown analyses (Brunesi & Parisi, 2017),
highlighting that seismic design according to Eurocode 8 (EN1998-1:2004) produced a
significant increase in vertical load capacity, ranging between 50% and 80%. Those results
confirmed previous numerical analyses that showed higher ultimate load factor of
earthquake-resistant buildings, with a mean overload factor greater than unity in case of
single-column-loss scenarios (Parisi & Augenti, 2012). Nonetheless, the case-study
building designed for earthquake resistance revealed an insufficient robustness under the
loss of a single column in a building façade with few bays. Accordingly, Li & Sasani (2015)
found that special frames designed according to American codes do not necessarily perform
better than their ordinary (i.e. non-seismic) counterparts in resisting progressive collapse.
This shows that the positive or negative impacts of earthquake-resistant design on
robustness are still a matter of research before general conclusions can be drawn. Hence,
seismic design not necessarily leads to an acceptable design for robustness.
Experimental tests have also been carried out to provide further evidence on this issue.
Sadek et al. (2011) and Lew et al. (2013) tested cast-in-place reinforced concrete sub-
assemblages, which consisted of two span beams, two exterior columns, and a central
column that was pulled down. Two types of specimen designed for different seismic
categories, namely, intermediate and special moment frames, were experimentally
studied. Yu and Tan (2013) tested two specimens with different seismic detailing. That
experimental programme was extended to four specimens with non-seismic and seismic
detailing (Yu & Tan, 2017) to evaluate the different behaviour of gravity-load-designed and
earthquake-resistant structures.
Both numerical and experimental studies on the assessment of structural resistance
against both earthquake actions and progressive collapse can promote innovation in design
and construction. Recent studies have evaluated the effectiveness of novel solutions in that
direction, which can further stimulate other studies aimed at reaching general conclusions

73
on the interaction between seismic and progressive collapse designs (Kim et al., 2011;
Feng et al., 2017; Lin et al., 2019a, 2019b; Lu et al., 2019; Quiel et al., 2019).
Based on static and dynamic nonlinear analyses, Kim et al. (2011) found that the combined
use of rotational friction dampers and high-strength tendons can significantly enhance both
the seismic performance and progressive collapse resistance of existing structures.
Feng et al. (2017) investigated new RC frame structures, proposing a novel kinked rebar
configuration to simultaneously improve earthquake resistance and robustness. The
location of kinked rebar in beams was optimized so that the progressive collapse resistance
was improved. From a physical standpoint, a kinked rebar has greater deformability as it
can be gradually straightened under tension. This produces a stair-stepped tensile
behaviour with a short elastic branch, followed by a plastic branch with very low hardening,
a third branch with very high hardening, and a final perfectly-plastic branch till failure. Two
yielding points are then identified at low and high stress/strain levels. Nonetheless, the
structural behaviour may be affected by several drawbacks, such as (i) the initial bending
capacity of the RC cross section lower than that associated with traditional steel
reinforcement, and (ii) possible shear failure when kinked rebar is located within the shear
span of beams.
Lin et al. (2019a, 2019b) remarked that considering seismic and progressive collapse
designs individually may produce an undesirable performance of the structure as well as
waste of construction materials. Those researchers proposed a novel design solution for
precast RC frame structures, which was validated via cyclic and progressive collapse tests,
the latter reproducing a middle-column removal scenario. Specifically, the case-study
structure was a multi-storey frame with precast RC beams and columns, which were
connected to each other by means of unbonded post-tensioning (PT) tendons, energy
dissipating steel angles, and shear plates. The proposed design solution was able to provide
the frame with important capacity features, such as large rotational capacity of beams,
slight damage, self-centring, and ease of repair.
Precast RC frame structures were also investigated by Quiel et al. (2019) who proposed
two variants of non-emulative beam-column connection for progressive collapse
resistance. The study focused on a ten-storey building with perimeter special moment
frames, which were subjected to ground-floor column removal. The proposed beam-column
connections consist of unbonded, high-strength steel PT bars, which pass through ducts in
the column and are anchored to the beams via bearing plates. PT bars act as structural
fuses after yielding, then maximising ductility. After that the design solution was validated
through full-scale pushdown testing, experimental results were incorporated into a
structural model of the frame system that was analysed under column removal. The
outcomes of nonlinear dynamic analysis showed that the structural system can arrest
progressive collapse under a single-column loss scenario.
Lu et al. (2019) proposed another design solution for composite steel-concrete frames
consisting of concrete-filled steel tube columns and steel I-beams, prestressed steel
strands, replaceable energy-dissipating components, and shear panels. The strands can
develop a self-centring capacity, hence minimising residual deformations and maximising
repairability. The energy-dissipating components were made of steel angles and rib
stiffeners. Based on experimental tests and finite element simulations, the proposed design
solution allowed the composite frame to develop better seismic and progressive collapse
performances compared to the traditional frame. Regarding the earthquake resistance, the
innovative frame demonstrated smaller residual deformations. At the same time, the
progressive collapse resistance was also improved through catenary action, increasing
rotational capacity of beam-column connections.
Although few studies have investigated the multi-hazard design for seismic resistance and
structural robustness, their outcomes clearly indicate some interesting chances to meet
multiple performance objectives, allowing the structure to develop different behavioural
modes depending on the type of actions they are subjected to. This promotes technological

74
innovation in the field that is aimed at increasing structural safety, resilience, and
sustainability.
Note that in the Eurocode system, seismic aspects are dealt with in EN 1998, supplemented
by country-dependent specifications is national annexes and treated by alternative limit
state formulations.

4.9 Fire design versus robustness


The fire action induces two main phenomena in the affected structure:
— Modification of the mechanical properties of the materials exposed to elevated
temperatures; and
— Redistribution of the existing internal stresses and development of new internal forces,
for example due to thermal expansion.
Obviously, these phenomena have to be accounted for when considering the fire design of
a structure. The consideration of the second one requires ensuring that the structures and,
in particular, key structural elements are able to sustain these extra forces.
When a fire develops in a structure, the temperature of the affected structural members is
increasing, inducing an elongation of the latter; so the development of axial compression
loads in these members is generally observed (when these members are axially restrained
at their extremities). Also, while the fire is developing, the material mechanical properties
are modified; in particular, their elastic strength and their young modulus are decreasing.
These modifications lead to the development of plastic zones in the structure but may also
provide these plastic zones with an additional deformation capacity in comparison to what
would be available at room temperature, leading to significant deformations and
displacements, in particular in the horizontal structural elements (i.e. the beams and/or
the slabs). Accordingly, even if the effect of the member elongation is governing at the
beginning of the fire with the evolution of axial compression loads, these axial loads start
to decrease at a certain time, i.e. when significant displacement appear, these can become
axial membrane tensile loads if the fire action and the associated elevation of temperature
are sufficient.
So, in most cases, a fire design requires ensuring that structural horizontal members are
able to sustain axial compression and tension loads which is also a requirement regularly
met when designing for robustness.
As an example, recent researches in field of fire design resulted in the development of
design recommendations to ensure the ability of a composite floor to develop membrane
forces and so to resist to fire actions (Vassart & Zhao, 2013). It seems obvious that the
application of these recommendations can be of help when considering, for instance, the
possibility of activating tying forces or when applying a column loss scenario in the
framework of a design for robustness. However, to date, it has not yet been demonstrated
that these rules are sufficient.
In addition, for redundant structures the fire performance cannot be evaluated by
considering the evolution of the thermal-induced damage at the local level – i.e. at the
cross-sectional or member level as addressed by design codes – but needs to be
investigated at the system level by taking into account the actual role played by the static
scheme in the time-variant stress redistribution process. With this regard, the margin of
safety may strongly depend on the prescribed fire scenario, and the most critical scenario
may be not associated with the maximum thermal load. Moreover, for a prescribed fire
scenario the structural performance is depending on both mechanical and thermal loading
history, and the most damaged structural configuration at the end of fire exposition is
generally not the most critical one (Biondini & Nero 2011). These aspects are crucial to
properly estimate thermo-mechanical damage effects and related consequences under fire
and establish suitable design rules to ensure structural robustness against fire events and
other interacting hazards.

75
Another approach which is also used in fire design is the compartmentation (i) to limit the
propagation of a fire but also (ii) to limit the propagation of the damages of a localized fire.
This last point requires to design the rest of the structure, i.e. the part of the structure not
directly affected by the localised fire, to sustain the additional forces associated to the fire
action and, in particular, the axial forces developing in the horizontal structural elements
affected by the fire. Again, the fact that specific parts of a structure are designed to resist
such loads can be seen as an added value when considering the limitation of the
propagation of a local damage in the framework of a design for robustness.
Considering these different points, it appears clearly that some structural requirements
associated to a fire design are concomitant with some associated to a design for
robustness. However, at this stage, it has not yet been demonstrated that the level of
requirements coincides, i.e. the fact of satisfying the requirements for fire design is
sufficient to ensure an appropriate level of robustness.
In a recent European project (Demonceau et al., 2013), it has been demonstrated that the
fact of considering both structural requirements, i.e. the ones from the fire design and the
ones from the design for robustness, in a combined design approach allows to guarantee
the requested level of safety while limiting the needs, for instance, in terms of fire
protection without any extra costs. However, this project was only covering a specific
structure typology, i.e. car parks, and further developments in this field are still required.
With respect to interacting hazards, in (Demonceau et al., 2013) a genuine example was
also presented of a coupled multi-hazard fire/robustness situation, where the local damage
is initiated by a localised fire, which can be treated within an extended robustness
assessment framework (Fang et al., 2013).
Note that in the Eurocode system, fire aspects are dealt with in parts 1-2 of EN1991 and
the difference material Eurocodes, supplemented by country-dependent specifications is
national annexes, and treated by alternative limit state formulations.

76
References
Adam, J., Parisi, F., Sagaseta, J., Lu, X. ‘Research and practice on progressive collapse and
robustness of building structures in the 21st century’, Engineering Structures, 173, 122-
149, 2018.
André, J., Canisius, G., Faber, M., Morton, J., ‘Background document – Robustness in the
Eurocodes – Project Team WG6.T1’, European Committee for Standardization (CEN), 2017,
89p.
ASCE, Causes of Silver Bridge Collapse Studied. Civil Engineering, American Society of Civil
Engineers, 38(12), 87, 1968.
Biondini, F., Camnasio, E., Palermo, A. ‘Lifetime seismic performance of concrete bridges
exposed to corrosion’, Structure and Infrastructure Engineering, Taylor & Francis, 10(7),
2014, 880-900.
Biondini, F., Frangopol, D.M., ‘Probabilistic limit analysis and lifetime prediction of concrete
structures’, Structure and Infrastructure Engineering, Taylor & Francis, 4(5), 399-412,
2008.
Biondini, F., Nero, A., ‘Cellular finite beam element for nonlinear analysis of concrete
structures under fire’, Journal of Structural Engineering, ASCE, 137(5), 543-558, 2011.
Biondini, F., Palermo, A., Toniolo, G., ’Seismic performance of concrete structures exposed
to corrosion: Case studies of low-rise precast buildings’, Structure and Infrastructure
Engineering, Taylor & Francis, 7(1–2), 109–119, 2011.
Biondini, F., ‘A Measure of Lifetime Structural Robustness’, ASCE/SEI Structures Congress
2009, Austin, TX, USA, April 30-May 2, 2009. In: Proceedings of the Structures Congress
2009, L. Griffis, T. Helwig, M. Waggoner, M. Hoit (Eds.), American Society of Civil
Engineers, USA, 2009.
Biondini, F., Frangopol, D.M., ‘Time-variant Robustness of Ageing Structures’, Chapter
6, Maintenance and Safety of Ageing Infrastructure, Y. Tsompanakis & D.M. Frangopol
(Eds.), Structures and Infrastructures Book Series, CRC Press, Taylor & Francis Group, 10,
163-200, 2014.
Biondini, F., Frangopol, D.M., ‘Life-cycle performance of deteriorating structural systems
under uncertainty: Review’, Journal of Structural Engineering, ASCE, 142(9), F4016001,
1-17, 2016.
Biondini, F., Frangopol, D.M., Restelli, S., ‘On Structural Robustness, Redundancy and
Static Indeterminacy’, ASCE Structures Congress 2008, Vancouver, B.C., Canada, 2008.
Biondini, F., Restelli, S., ‘Damage Propagation and Structural Robustness’, First
International Symposium on Life-Cycle Structural Engineering (IALCCE'08), Varenna, Italy,
June 10-14, 2008. In Life-Cycle Civil Engineering, F. Biondini and D.M. Frangopol (Eds.),
CRC Press, Taylor and Francis Group, 131-136.
Brunesi, E., Nascimbene, R., Parisi, F., Augenti, N. ‘Progressive collapse fragility of
reinforced concrete framed structures through incremental dynamic analysis’, Engineering
Structures, 104, 65-79, 2015.
Brunesi, E., Parisi, F., ‘Progressive collapse fragility models of European reinforced concrete
framed buildings based on pushdown analysis’, Engineering Structures, 152, 579-596,
2017.
CEN (2021) EN pr1990: Eurocode 0 - Basis of structural design (prEN 1990:2020). Brussels
(Belgium).
Corley, W.G. ‘Lesson learned on improving resistance of buildings to terrorist attacks’,
Journal of Performance of Constructed Facilities, ASCE, 18(2), 68-78, 2004.
Demonceau et al., ‘Robustness of car parks against localised fire’, European Commission –
RFCS publications EUR 25864, ISBN 978-92-79-28956-9, 2013.

77
Dusenberry, D.O., Juneja, G., ‘Review of existing guidelines and provisions related to
progressive collapse’, Workshop on Prevention of Progressive Collapse, National Institute
of Building Sciences (NIST), 2002.
Ellingwood, B.R., ‘Risk-informed condition assessment of civil infrastructure: state of
practice and research issues’, Structure and Infrastructure Engineering, 1(1), 7–18, 2005.
Ellingwood, B.R., Mitigating risk from abnormal loads and progressive collapse. Journal of
Performance of Constructed Facilities, ASCE, 20(4), 315-323, 2006.
Ellingwood, B., Leyendecker E.V., ‘Approaches for design against progressive collapse’,
ASCE: J. Struct. Div; 104(3): 413–423, 1978.
EN 1991-1-7, ‘Eurocode 1 – Actions on structures, Part 1-7: General actions – accidental
actions’, Comité European de Normalization (CEN), 2006.
EN 1998-1. ‘Eurocode 8 – Design of structures for earthquake resistance, Part 1: General
rules, seismic actions and rules for buildings, Comité European de Normalization (CEN),
2004.
Fang, C., Izzuddin, B.A., Elghazouli, A.Y., Nethercot, D.A., ‘Simplified energy-based
robustness assessment for steel-composite car parks under vehicle fire’, Engineering
Structures, 49, 719–732, 2013.
Feng, P., Qiang, H., Qin, W., Gao, M. ‘A novel kinked rebar configuration for simultaneously
improving the seismic performance and progressive collapse resistance of RC frame
structures’, Engineering Structures, 147, 752–767, 2017.
fib MC 2020 draft
fib (2006) fib Bulletin 34: Model code for service life design of concrete structures,
Fédération Internationale du Béton (fib)
Gross, J. L., McGuire, W., ‘Progressive collapse resistant design’, Journal of Structural
Engineering, ASCE, 109(1), 1–15, 1983.
Gurley, C. Progressive collapse and earthquake resistance. Practice Periodical on Structural
Design and Construction, 13(1), 19-23, 2008.
Hayes, J.R., Woodson, S.C., Pekelnicky, R.G., Poland, C.D., Corley, W.G., Sozen, M. ‘Can
strengthening for earthquake improve blast and progressive collapse resistance?’ Journal
of Structural Engineering, ASCE, 131(8), 1157-1177, 2005.
Kim, J., Choi, H., Min, K.W. ‘Use of rotational friction dampers to enhance seismic and
progressive collapse resisting capacity of structures’, The Design of Tall and Special
Buildings, 20, 515-537, 2011.
Lalkovski, N., Starossek, U. ‘Vertical building collapse triggered by loss of all columns in
the ground story − Last line of defense’, International Journal of Steel Structures, 16(2),
395-410, 2016.
Lew, H.S., Main, J.A., Robert, S.D., Sadek, F., Chiarito, V.P. ‘Performance of steel moment
connections under a column removal scenario. I: Experiments’, Journal of Structural
Engineering, ASCE, 139(1), 98-107, 2013.
Li, M., Sasani, M., ‘Integrity and progressive collapse resistance of RC structures with
ordinary and special moment frames’, Engineering Structures, 95, 71-79, 2015.
Li, Y., Ahuja, A., Padgett, J.E., ‘Review of methods to assess, design for, and mitigate
multiple hazards’, Journal of Performance of Constructed Facilities, ASCE, 26(1), 104–117,
2011.
Lin, K., Lu, X., Li, Y., Guand, H. ‘Experimental study of a novel multi-hazard resistant
prefabricated concrete frame structure’, Soil Dynamics and Earthquake Engineering, 119,
390-407, 2019a.

78
Lin, K., Lu, X., Li, Y., Zhuo, W., Guan, H. ‘A novel structural detailing for the improvement
of seismic and progressive collapse performances of RC frames’, Earthquake Engineering
and Structural Dynamics, 48, 1451-1470, 2019b.
Lin, K.Q., Li, Y., Lu, X., Guan, H., ‘Effects of seismic and progressive collapse designs on
the vulnerability of RC frame structures’, Journal of Performance of Constructed Facilities,
ASCE, 31(1), 04016079, 2016.
Livingston, E., Sasani, M., Bazan, M., Sagiroglu, S., ‘Progressive collapse resistance of RC
beams’, Engineering Structures, 95, 61-70, 2015.
Lu, X., Zhang, L., Lin, K., Li, Y. ‘Improvement to composite frame systems for seismic and
progressive collapse resistance’, Engineering Structures, 186, 227-242, 2019.
NTSB, Collapse of Suspended span of Route 95 Highway Bridge over the Mianus River
Greenwich, Connecticut June 28, 1983. Highway Accident Report, National Transportation
Safety Board, Washington D.C., 1983.
Okasha, N.M. & Frangopol, D.M., Time-variant redundancy of structural systems, Structure
and Infrastructure Engineering, 6, 279-301, 2010.
Parisi, F., Augenti, N. ‘Influence of seismic design criteria on blast resistance of RC framed
buildings: A case study’, Engineering Structures, 44, 78–93, 2012.
Park, J., Kim, J. ‘Fragility analysis of steel moment frames with various seismic connections
subjected to sudden loss of a column’, Engineering Structures, 32(6), 1547-1555, 2010.
Pekau, O.A, Cui, Y. ‘Progressive collapse simulation of precast panel shear walls during
earthquakes’, Computers and Structures, 84, 400-412, 2006.
Quiel, S.E., Naito, C.J., Fallon, C.T. ‘A non-emulative moment connection for progressive
collapse resistance in precast concrete building frames’, Engineering Structures, 179, 174-
188, 2019.
Sadek, F., Main, J.A., Lew, H.S., Bao, Y. ‘Testing and analysis of steel and concrete beam-
column assemblies under a column removal scenario’, Journal of Structural Engineering,
ASCE, 137(9), 881-892, 2011.
Starossek, U., ‘Disproportionate Collapse: A Pragmatic Approach’, proceedings of the
Institution of Civil Engineers – Structures and Buildings, 160 (6): 317-325, 2007.
https://doi.org/10.1680/stbu.2007.160.6.317
Starossek, U., ‘Progressive Collapse of Structures’, Vol. 153, Thomas Telford, London,
2009.
Starossek, U., Haberland, M., ‘Robustness of Structures’, International Journal of Lifecycle
Performance Engineering, 1, 3-21, 2012.
Starossek, U., Haberland, M., ‘Measures of structural robustness – Requirements &
applications’, ASCE Structures Congress 2008, Vancouver, B.C., Canada, 2008.
Tanner P, Hingorani R., ‘Robustness: a practitioner’s perspective’, IABSE Symposium
Guimaraes, Towards a Resilient Built Environment - Risk and Asset Management,
Guimaraes, 2019
Tsai, M.-H., Lin, B.-H. ‘Investigation of progressive collapse resistance and inelastic
response for an earthquake-resistant RC building subjected to column failure’, Engineering
Structures, 30(12), 3619-3628, 2008.
Vassart, O., Zhao, B., ‘Membrane action of composite structures in case of fire’, publication
N° 132 from ECCS, 2013.
Woliński, S., ‘Defining of the structural robustness’, Bulletin of the Polish Academy of
Sciences Technical Sciences 61, 1:137-144, 2013.

79
Xu, G., Ellingwood, B.R., ‘Probabilistic robustness assessment of pre-Northridge steel
moment resisting frames’, Journal of Structural Engineering, ASCE, 137(9), 925-934,
2011.
Yu, J., Tan, K.H. ‘Experimental and numerical investigation on progressive collapse
resistance of reinforced concrete beam column sub-assemblages’, Engineering Structures,
55, 90-106, 2013.
Yu, J., Tan, K.H. ‘Special detailing techniques to improve structural resistance against
progressive collapse’, Journal of Structural Engineering, ASCE, 140(3), 1-15, 2014.
Yu, J., Tan, K.H. ‘Structural behaviour of reinforced concrete frames subjected to
progressive collapse’, ACI Structural Journal, 114(1), 63-74, 2017.
Zhu, B., Frangopol, D.M., ‘Reliability, redundancy and risk as performance indicators of
structural system during their life-cycle’. Engineering Structures, 41, 34-49, 2012.

80
5 Quantitative measures of structural robustness
Structural robustness evaluations should provide objective quantifications to establish
design verification criteria and ranking design solution alternatives. Furthermore, a
quantitative measure of system robustness should allow to prioritize maintenance and
repair interventions on existing structures. However, despite the significant efforts made
to develop robustness quantification criteria and procedures considering sudden and
continuous damage under time-dependent exposure scenarios, there are no widely
accepted criteria in literature for a definition and quantitative measure of robustness.
Furthermore, civil engineering standards and design codes do not provide methodologies
or specifications for robustness quantification. Moreover, several metrics proposed in
literature are formulated on the basis of concepts that may be related to structural
robustness, like risk, vulnerability, redundancy, residual strength, and damage tolerance,
but that in general do not provide robustness measures. However, at the basis of the
proposed approaches lies the same general idea: the comparison of the intact and
damaged structure by means of structural performance indicators. This procedure can be
formulated in a deterministic or probabilistic approach, accounting for uncertainties and
consequences including risk quantification. Moreover, robustness evaluations should not
be restricted to accidental actions and abnormal loadings. Ageing and deterioration
processes, the effects of design, construction and maintenance errors may also lead to
disproportionate effects.
In the following, several robustness indicators under both linear elastic and nonlinear
behaviour are briefly introduced. Subsequently, structural analysis methods to be used for
quantification of performance indicators under damage are presented. Finally, criteria for
robustness quantification are formulated.

5.1 Structural performance indicators for robustness assessment


Structural robustness is associated with the ability of the system to avoid consequences
that are disproportionate with respect to the extent of the triggering damaging event.
According to this definition, both the loss of performance when damage occurs and the
amount of damage needs to be considered to achieve meaningful robustness evaluations.
Performance indicators formulated as the ratio of a performance parameter of the intact
and damaged system are hence used as state variables (Frangopol & Curley 1987, Biondini
& Restelli 2008).
The selection of suitable performance indicators represents a critical task and should be
based on both the limit state condition and exposure scenario to be investigated. Moreover,
in order to effectively describe the effects of damage on the system, structural performance
indicators should be able to capture the role played by the damaged members and identify
failure conditions and damage propagation.
Strength and ductility, as well as other performance indicators of nonlinear behaviour, may
be used in robustness evaluations associated with damage induced by severe loadings,
such as explosions or impacts. However, performance indicators of the serviceability
conditions under linear behaviour, such as elastic stiffness and first yielding, may become
of major importance in life-cycle robustness evaluations associated with aging of
structures. In addition, it has been noted that the assumption of linear behaviour can be
successfully used in design of robust structures (Powell 2009).
Robustness indicators can be formulated in deterministic terms. However, civil engineering
systems are usually characterized by significant uncertainties related to structural
modelling, exposure scenario, loading conditions, and failure consequences. Probability-
based or reliability-based formulations can be adopted if uncertainties are incorporated in
the robustness indicators. Moreover, risk-based robustness indicators can be formulated
based on a suitable quantification of the failure consequences.
The robustness indicators presented in the following can effectively be used to relatively
compare the system robustness with respect to different damage scenarios and exposures.

81
However, such indicators need to be complemented with the amount of damage to quantify
in absolute terms the disproportion of damage effects and structural robustness.

5.1.1 Deterministic indicators


The effectiveness of several dimensionless performance indicators 0≤≤1 related to the
structural behaviour of linear systems in evaluating structural robustness is investigated in
Biondini & Restelli (2008). Indicators associated with the properties of the structural
system only, such as determinant, trace, and condition number of the overall stiffness
matrix and first natural vibration period, and indicators depending also on the loading
scenario, including stored energy and displacements, are considered. It is found that
determinant, trace, and condition number of the stiffness matrix are not suitable to
effectively describe the effects of selected damage scenarios on the structural performance
and the following indicators are recommended:
Tn 0
T  Tn  2 maxi (K 1M ) (5-1)
Tn1 i

s0
s  s  s  K 1f (5-2)
s1

0 1 1
    sT Ks  sT f (5-3)
1 2 2
where Tn is the first natural vibration period associated with the stiffness matrix K and
mass matrix M, i(A) denotes the ith eigenvalue of a square matrix A, s is a displacement
vector, f is a load vector,  is the stored energy,  denotes the Euclidean scalar norm,
and the subscripts “0” and “1” refer to the intact and damaged states, respectively.
Starossek & Haberland (2011) also proposed a stiffness-based robustness indicator based
on the minimum determinant of the stiffness matrix after removing a single structural
component or connection j, i.e. for the worst-case damage scenario, as follows:
det(𝐊j )
𝑅𝑠 = min (5-4)
𝑗 det(𝐊0 )

Also in this case, the authors reported that the expressiveness of a stiffness-based
robustness indicator is not sufficient and recommend the use of the following energy-based
indicator:
𝐸𝑟,𝑗
𝑅𝑒 = 1 − max (5-5)
𝑗 𝐸𝑓,𝑘

where 𝐸𝑟,𝑗 is the energy released during initial failure of structural element 𝑗 and
contributing to damaging a subsequently affected element 𝑘, and 𝐸𝑓,𝑘 is the energy required
for the failure of the subsequentially element 𝑘.
For robustness evaluations it is also of interest to define indicators able to simultaneously
account for the structural performance of both the intact and damaged system. To this
purpose, vectors of nodal forces equivalent to the effects of damage, defined as backward
or forward pseudo-loads, are also considered in Biondini & Restelli (2008). The concept of
pseudo-load is qualitatively explained in Figure 5-1 based on the linear equilibrium
equations of both the intact and damaged systems (Figure 5-1.a):
K 0s 0  f 0 K1s1  f1 (5-6)

The displacement vector of the intact system s0 can be related to the displacement vector
of the damaged system s1 and related stored energy variation 1 as follows:

s 0  s1  K 11fˆ1  K 11 (f1  fˆ1 ) (5-7a)

82
fˆ1  (K1  K 0 )s0  (f1  f0 )  Ks0  f (5-7b)

ˆ  1 s T f  1 s T (f  fˆ )   1 s T ( fˆ  f )
 0   0   (5-7c)
1 0 0 0 1 1 0 1
2 2 2
where f̂1 is a vector of nodal forces equivalent to the effects of repair, or backward pseudo-
load vector, and ̂1 is the stored energy associated with the damaged system after the
application of the backward pseudo-loads (area OP0 P̂1 in Figure 5-1.a for the case f=0).
The vector f̂1 represents the additional nodal forces that must be applied to the damaged
system to achieve the nodal displacements of the intact system (Figure 5-1.b).
In a dual way, the displacement vector of the damaged system s1 can be related to the
displacement vector of the intact system s0 and related stored energy variation 0 as
follows:

s 1  s 0  K 01fˆ0  K 01 (f 0  fˆ0 ) (5-8a)

fˆ0  (K1  K 0 )s1  (f1  f0 )  Ks1  f (5-8b)

ˆ    1 s T (f  fˆ )  1 s T f  1 s T ( fˆ  f )
 1   (5-8c)
0 1 1 0 0 1 1 1 0
2 2 2
where f̂ 0 is a vector of nodal forces equivalent to the effects of damage, or forward pseudo-
load vector, and ̂ 0 is the stored energy associated with the intact system after the
application of the forward pseudo-loads (area OP̂0 P1 in Figure 5-1.a for the case f=0).
The vector f̂ 0 represents the additional nodal forces that must be applied to the intact
system to achieve the nodal displacements of the damaged system (Figure 5-1.c).

Figure 5-1. Force f=f0=f1 versus displacement s of a truss system in the intact state and after
elimination of one member. (a) Force-displacement diagrams. (b) Backward pseudo-loads (effects of
repair). (c) Forward pseudo-loads (effects of damage).

Backward Forward
Pseudo-Loads Pseudo-Loads

 0 = Area OP0 P̂1  1 = Area OP̂0 P1

(a) (b) (c)


Source: Biondini & Restelli, 2008.

It is found that backward pseudo-loads exhibit little sensitivity to damage, particularly


under extensive damage. Contrary, forward pseudo-loads can effectively be used for
robustness evaluations and the use of the following robustness indicator is preferred:
1
1  1  (5-9)
ˆ
0

83
Indicators associated with first failure and structural collapse have been also investigated
by several authors. The effect of damage on structural performance is quantified in terms
of structural redundancy by Frangopol and Curley (1987) based on residual load-carrying
capacity as follows:
𝐿𝑖𝑛𝑡𝑎𝑐𝑡
𝑅𝐿 = (5-10)
𝐿𝑖𝑛𝑡𝑎𝑐𝑡 −𝐿𝑑𝑎𝑚𝑎𝑔𝑒𝑑

where 𝐿𝑖𝑛𝑡𝑎𝑐𝑡 and 𝐿𝑑𝑎𝑚𝑎𝑔𝑒𝑑 are the collapse loads for the intact and damaged system,
respectively. This measure is effective to investigate the reserve of load capacity after
damage occurs and the importance of individual structural components.
Ghosn & Moses (1998) formulated a redundancy factor for highway bridges considering
the load factors associated with ultimate (𝑢), serviceability (𝑠), and damage (𝑑) limit states.
The system reserve factor 𝑅𝐿𝐹 is defined as follows:
𝐿𝐹𝑢,𝑠,𝑑
𝑅𝑢,𝑠,𝑑 = (5-11)
𝐿𝐹1

where 𝐿𝐹𝑢,𝑠,𝑑 is the load factor which exceeds a specific limit state (u, s, d) and 𝐿𝐹1 is the load
multiplier of failed member. A unit value of the reserve ratio corresponds to a non-
redundant system with respect to the failure of the analysed member.
Wisniewski et al. (2006) developed a similar method to evaluate robustness of railway
bridges. According to this approach, structural robustness is defined as the ability of the
system to continue carrying loads after a member failure and can be quantified using
redundancy ratios which compare system and member capacities at the serviceability or
ultimate limit state.
Along similar research lines, Maes et al. (2006) proposes the use of a Reserve Strength
Ratio (𝑅𝑆𝑅) associated with the load-carrying capacity evaluated in the damaged (𝑅𝑆𝑅𝑖 )
and undamaged (𝑅𝑆𝑅0 ) state as follows:
𝑅𝑆𝑅𝑖
𝑅1 = min (5-12)
𝑖 𝑅𝑆𝑅0

where minimization allows to capture the worst case scenario associated with damage of
the i-th individual structural component.
Indicators associated with first failure and sequential failures up to structural collapse can
be found also in Biondini & Frangopol (2014, 2017). Denoting   0 a scalar load multiplier,
the limit states associated to the occurrence of a series of sequential failures k1,2,… can
be identified by the corresponding failure load multiplier k. The ability of the system to
redistribute the load after the failure ki up to the failure kj depends on the reserve load
carrying capacity associated to the failure load multipliers i=i and j=j and the following
quantity can be assumed as a measure of redundancy between subsequent failures:
𝜆𝑗 −𝜆𝑖
Λ𝑖𝑗 = (5-13)
𝜆𝑗

This factor can assume values in the range [0;1]. It is zero when there is no reserve of
load capacity between the failures i and j (i=j), and tends to unity when the failure load
capacity i is negligible with respect to j (i<<j). It is worth noting that this definition
incorporates the classical measure of redundancy associated with the ability of the system
to redistribute the load after the occurrence of the first local failure, reached for i=1, up
to structural collapse, reached for a collapse load multiplier j=c:
𝜆𝑐 −𝜆1
Λ= (5-14)
𝜆𝑐

This redundancy measure is further generalized for deteriorating systems to introduce the
concept of failure times and elapsed times between subsequent failures (Biondini 2012).
Progressive collapse of systems is studied by analogy with fast fracture in metals by Smith
(2006). The energy released in structural component-loss damage scenarios is compared
with the energy absorbed by the damaged members, like in metal cracks propagation. If

84
the energy released is greater than the energy adsorbed, progressive collapse will occur.
The proposed methodology also allows to identify critical sequence of damaged members
by sequentially removing damaged elements and solving a minimisation process on the
damage energy.
André et al. (2015) point out some limitations in the approach proposed by Smith (2006)
and proposes an energy-based robustness indicator defined as follows:
𝐷𝑢𝑐 −𝐷1𝑠𝑡 𝑓𝑎𝑖𝑙𝑢𝑟𝑒
𝐼𝑅 (𝐴𝐿 |𝐻) = (5-15)
𝐷𝑐 −𝐷1𝑠𝑡 𝑓𝑎𝑖𝑙𝑢𝑟𝑒

where 𝐴𝐿 is the leading action, 𝐻 is a set of hazard scenarios and 𝐷1𝑠𝑡𝑓𝑎𝑖𝑙𝑢𝑟𝑒 , 𝐷𝑢𝑐 , and 𝐷𝑐 , are
the damage energies associated respectively with first failure, “unavoidable collapse” state,
and the structural collapse. This indicator allows to capture the role of damage propagation
on the value of “unavoidable collapse” energy state and how the latter compares with the
potential maximum energy state of the system.
Robustness assessment should involve a quantification of the direct and indirect failure
consequences (which is inherently done e.g. in case of applying a risk-based robustness
indicator). To this purpose it is important to consider that structural collapse may have
different consequences. For example, the global collapse of a whole structural system
should be considered more important than the local collapse of a single member or a
portion of the structure (Figure 5-2). Despite the evaluation of failure consequences is a
challenging task and a reliable estimation is often not straightforward, a relatively simple
importance measure of structural failure could be provided by the following structural
integrity index (Biondini & Restelli 2008):
V1
V  (5-16)
V0
where V1 is the portion of structural volume V0 which remains intact after damage. Failed
members involved in a collapse mechanism can be identified based on the eigenvectors si
of the stiffness matrix K associated with the eigenvalues i(K)=0.

Figure 5-2. Types of failure (adapted from Biondini & Restelli 2008): (a) local and (b) global
collapse.
Damage

(a)
Damage

(b)

Source: Biondini & Restelli, 2008.

85
5.1.2 Probability-based and reliability-based indicators
The robustness indicators previously introduced can be formulated in a probabilistic context
to account for the uncertainties related to structural model, exposure scenario, loading
conditions, and failure consequences, among others. However, indicators explicitly related
to quantification of probability of failure or reliability index have been proposed to directly
formulate robustness indicators in probabilistic terms.
Frangopol & Curley (1987) proposed a redundancy measure based on the reliability index
𝛽 computed with respect to serviceability or ultimate limit state conditions as follows:
𝛽𝑖𝑛𝑡𝑎𝑐𝑡 (5-17)
𝑅𝛽 =
𝛽𝑖𝑛𝑡𝑎𝑐𝑡 − 𝛽𝑑𝑎𝑚𝑎𝑔𝑒𝑑

where 𝛽𝑖𝑛𝑡𝑎𝑐𝑡 and 𝛽𝑑𝑎𝑚𝑎𝑔𝑒𝑑 are the reliability indices of the intact and damaged systems,
respectively.
A measure of the effects of damage evaluated in probabilistic terms was proposed by Lind
(1995) based on the concept of structural vulnerability, defined as the insensitivity of the
system to damage, as follows:
𝑃(𝑟𝑑 ) (5-18)
𝑉=
𝑃(𝑟0 )

where 𝑃(𝑟) is the probability of failure in the damaged (𝑟𝑑 ) and intact (𝑟0 ) state.
Maes et al. (2006) proposes a similar approach by a minimisation aimed at identifying the
worst case scenario associated with damage of the i-th individual structural component:
𝑃𝑠0 (5-19)
𝑅𝑃 = min
𝑖 𝑃𝑠𝑖

where 𝑃𝑠0 and 𝑃𝑠𝑖 are the probabilities of failure in the intact and damaged states,
respectively.

5.1.3 Risk-based indicators


Risk-based robustness assessment incorporates the quantification of failure consequences.
Baker et al. (2008) developed a risk-based approach where consequences are separated
into two contributions: Direct consequences associated to the damage of elements directly
affected by the hazardous event, and indirect consequences associated with the
subsequent partial or total system failure. Risk is computed by the product of probability
of occurrence of disproportionate collapse and corresponding consequences. The
quantification of consequences may also include both monetary and human losses.
Considering that a lower amount of indirect risk indicates a more robust system, the
following indicator is proposed:
𝑅𝐷𝑖𝑟 (5-20)
𝐼𝑅 =
𝑅𝐷𝑖𝑟 + 𝑅𝐼𝑛𝑑

where 𝑅𝐷𝑖𝑟 and 𝑅𝐼𝑛𝑑 are the direct and indirect risks, respectively, associated with a specific
exposure scenario. If the system failure does not involve indirect consequences, then 𝐼𝑅 =
1. On the other hand, 𝐼𝑅 = 0 if direct consequences are negligible compared with indirect
consequences. This approach can be extended to account for multiple exposure scenarios:
∑𝑖 𝑅𝐷𝑖𝑟 (5-21)
𝐼𝑅 =
∑𝑖 𝑅𝐷𝑖𝑟 + ∑𝑗 𝑅𝐼𝑛𝑑

where the risks for each scenario i need to be computed.

86
5.2 Structural analysis methods for robustness assessment

5.2.1 General
Quantification of progressive collapse resistance and robustness relies upon the analysis of
structural response to certain design scenarios, which should be defined according to
building codes, authorities, and stakeholders. In line with Sect. 4.4, each scenario consists
of either single or multiple damaging events/conditions that may occur in the structure’s
lifetime, producing the need to assess possible disproportionate consequences. In threat-
dependent design/assessment approaches, scenarios are delineated in the form of
specified accidental actions (e.g. fire, blast, impact or notional loads) and/or deterioration
processes (i.e. corrosion, fatigue, among others). These threats can be modelled according
to technical documents (e.g. building codes, standards, guidelines, and reports), scientific
literature, or ad-hoc investigations committed for the structure under study (as in the case
of, e.g., critical infrastructure). In threat-independent approaches, scenarios are defined
in terms of notional damage applied to single or a few components of the structure to
assess robustness and/or investigate progressive collapse (Biondini and Restelli, 2008;
Parisi and Scalvenzi, 2020). The latter is thus evaluated as the insensitivity of the structure
to such scenarios (see e.g. Sect. 2.1.3). Both threat-dependent and threat-independent
approaches are usually included in the general category of direct design methods, which
are aimed at explicitly evaluating the performance of the structural system to redistribute
loads in case of local damage and to avoid disproportionate collapse.
Several codes and guidelines include two types of direct design/assessment approaches,
namely the key element method and alternative load path (ALP) method (see e.g. Sects.
3.1, 3.2, 4.5.2 and 4.5.5). The former method focuses on the capacity of single
components to withstand abnormal loads, which for instance is a major situation for
peripheral columns of frame buildings due to their high exposure. Hence, the
implementation of the key element method requires modelling and analysis of selected
structural components. By contrast, the ALP method is based on the response analysis of
the structure under either specified abnormal actions or notional damage applied to some
components (e.g. Parisi and Augenti, 2012).
Regardless of the approach selected for robustness quantification and/or progressive
collapse analysis, the numerical analysis of the structure under damage scenarios is a key
issue to address and is discussed in the following sub-sections. In this respect, it should
be noted that robustness quantification is different from progressive collapse analysis.
Robustness analysis is aimed at comparing the loss of performance with the amount of
damage to evaluate the ability to avoid disproportionate consequences under prescribed
damage scenarios. Progressive collapse analysis investigates how damage will propapate
leading to structural collapse. Linear and limit analysis methods allow to investigate the
system performance in the elastic stage and at collapse and can be effectively used for
robustness quantification (Biondini & Restelli 2008, Biondini 2009, Powell 2009).
Conversely, progressive collapse analysis usually requires nonlinear analysis methods
because it is aimed at evaluating the propagation of damage throughout the structure.
Based on such considerations, Sects. 5.2.4, 5.2.5 and 5.2.6 describe different methods of
structural analysis according to an increasing level of complexity, moving from Level I
methods (involving linear and limit analyses), through Level II methods (including
nonlinear analysis procedures), and to Level III methods (i.e. probabilistic analyses). Such
an organization of this section is in line with most of building codes and guidelines at both
national and international levels, which allow several methods for response analysis of
structural systems, i.e. linear or nonlinear, static or dynamic analyses. In some cases,
easier calculation methods for structural modelling and response analysis can be used, i.e.,
avoiding more complex computations through numerical approaches.
Figure 5-3 shows a general overview of the most typical methods for structural response
analysis and robustness quantification under either specified accidental actions (threat-
dependent approaches; option A) or notional damage (threat-independent approaches;
option B). In case the structure is designed or assessed for specified accidental actions,

87
one can choose between key element design (option A1) or ALP analysis (option A2).
Conversely, design and assessment for notional damage can be followed by the application
of the ALP method (option B1) or directly lead to a basic framework for structural analysis
(option B2). Whichever option is selected, the structural engineer has to choose
appropriate methods for structural analysis. Sect. 5.2.2 briefly describe the following
methods: finite element method; discrete element method; applied element method; or
cohesive element method. If option A1 is selected, then numerical modelling focuses on
single components and can be carried out according to several computational strategies,
as discussed in Sect. 5.2.3. If option A2 is selected, the ALP method starts with numerical
analysis of single structural components, then moving to the analysis of the pre-damaged
structural system via one of the methods described in Sects. 5.2.4, 5.2.5 and 5.2.6. In
case of design/assessment for notional damage (option B), the response analysis of the
structural system is directly performed without analysing individual components.

Figure 5-3. Overview of structural analysis methods for progressive collapse analysis and
robustness quantification.

5.2.2 Numerical modelling strategies


The progress in computer technology and structural software has given rise to a number
of computational strategies that allow researchers and practitioners to assess complex
phenomena involving material nonlinearities, fragmentation, impact, large deformations,
and dynamics. Nonetheless, the level of sophistication should be accurately managed
depending on the type of structure, design scenarios and their potential consequences. For
instance, in progressive collapse simulation, combining the outputs of linear static response
analysis and rigid-plastic (or limit equilibrium) analysis can provide a computationally
efficient support to robustness quantification, according to indicators presented in Sect.
5.1. By contrast, more advanced, nonlinear static/dynamic analysis methods – which

88
require high computational expertise – are needed particularly when more complex cases
(e.g. the assessment of existing structures) must be investigated.
According to recent literature reviews (El-Tawil et al., 2014; Adam et al., 2018; Kunnath
et al., 2018), there are several numerical modelling strategies that have been successfully
validated through experimental tests at different structural scales (i.e. component, sub-
system, system). This allows realistic simulations of structural behaviour, saving cost and
time for further experimental testing.
Most of modelling strategies are based on the finite element method (FEM), which is
implemented with various degrees of approximation and complexity. FEM allows several
2D and 3D structural models to be developed, such as micro-models, macro-models, and
hybrid models. It is also noted that FEM allows almost any type of structural response
analysis, i.e. linear or nonlinear, static or dynamic, implicit or explicit, depending on the
expected accuracy level. Micro-modelling approaches are feasible for relatively small-scale
problems related to single structural components or sub-systems, whereas macro-models
are used to analyse whole structures. In the framework of micro-models, solid (brick)
elements can be used for detailed geometric representations and structural simulations
where different materials (e.g. concrete, steel rebar, welds, bolts) and their mutual
interaction (e.g. steel-concrete bond) is explicitly considered. Studies based on micro-
modelling FEM procedures were carried out by, amongst others, the following researchers:
Sadek et al. (2011) and Guo et al. (2013) for steel and composite structures; Shi et al.
(2011), Li et al. (2016) and Pham et al. (2017) for cast-in-place RC structures; and
Elsanadedy et al. (2017) for precast RC structures. Macro-models consist of beam/shell
elements, so their lower computational cost allows the progressive collapse analysis of
entire structures under either specified abnormal actions and notional scenarios (e.g.
Izzuddin et al., 2008; Fu et al., 2011; Parisi and Augenti, 2012; Kazemi-Moghaddam and
Sasani, 2015; Lu et al., 2017; Feng et al., 2021). In that context, beam-type elements are
normally used to model columns and beams, sometimes according to fibre-based modelling
approaches (e.g. Brunesi and Nascimbene, 2014; Mucedero et al., 2020). Effective beam
finite element models have been also proposed to incorporate the effects of aging and
deterioration processes, such as reinforcement corrosion in concrete structures under static
and seismic loadings (Biondini et al. 2013, Biondini & Vergani 2015). Shell elements are
utilised to simulate floor slabs and thin-walled steel components. The relative simplicity of
macro-models makes them the most used computational technique in engineering
computations for progressive collapse analysis and robustness assessment. Nonetheless,
higher complexity levels may arise from numerical simulation of beam-column and slab-
beam joints, particularly in existing RC frame structures not designed for earthquake
resistance (see e.g. Kunnath et al., 2018), precast RC frame structures (e.g. Ravasini et
al., 2021), as well as steel and composite structures. Those parts of structures mobilise
the interaction between large-deformation resisting mechanisms such as catenary action
of beams, membrane action of slabs, and axial-shear-flexure interaction. Beam-column
macro-models consisting of spring and rigid elements that are connected to beam/shell
elements were proposed by, e.g., Bao et al. (2008), Sadek et al. (2008), Liu et al. (2010),
Khandelwal and El-Tawil (2011), Jahromi et al. (2013), and Sun et al. (2015). In some
studies, different types of finite elements were combined with each other to capture both
local and global behavioural features of structures, delineating hybrid (or multi-scale)
models (e.g. Alashker et al., 2011, El Hajj Diab et al., 2021, 2022). Such models
significantly reduce the number of elements and computational cost, even though
kinematic compatibility between different elements is a critical issue to take into account.
In recent years, progressive collapse simulations have been carried out using the discrete
element method (DEM), which can be also combined with FEM to obtain accurate results
in a computationally efficient manner. DEM was originally developed to model granular and
discontinuous materials particularly in rock mechanics, but it was recently extended to
masonry structures and other types of engineering facilities. The interaction between
discrete elements is modelled through contact relationships, allowing the dynamic
behaviour of the material at the macro-scale to be derived from normal, tangent, rolling
and twisting motion of the interacting elements. The computational cost of DEM-based

89
numerical models is usually high, especially in case of structural sub-systems and systems
(e.g. Pekau and Cui, 2006; Masoero et al., 2010; Gu et al., 2014), but the use of discrete
elements coupled with finite elements can allow successful simulations accounting for
material fragmentation and impact (e.g. Lu et al., 2009).
Besides FEM and DEM, some advanced software packages have been developed by
implementing the applied element method (AEM) originally proposed by Tagel-Din and
Meguro (2000). The use of AEM is rapidly increasing for high-fidelity, progressive collapse
simulations of structural sub-systems and entire structures. This numerical method is
based on the modelling of the structure as an assembly of relatively small elements, which
are connected to each other at their contact (boundary) points through normal and shear
springs. The presence of discontinuities between applied elements allows realistic
simulations of cracking, separation, and collision of structural components up to collapse.
Latest developments in high-performance computing nowadays allow a moderate level of
computational cost for progressive collapse simulation, with successful application to both
structural sub-assemblages and entire structures (e.g. Dinu et al., 2016; Salem and Helmy,
2014; Salem et al., 2016; Khalil, 2012).
Another modelling strategy for progressive collapse analysis makes use of the cohesive
element method (CEM), which is widely effective for fracture mechanics problems. CEM
essentially smears a finite, potential damage zone by means of a zero-thickness nonlinear
element, while modelling undamaged parts via linear elastic elements. Nonlinear cohesive
elements are thus used to lump nonlinear behaviour of structural components where
damage is expected to occur during progressive collapse (Keyvani and Sasani, 2015). Le
and Xue (2014) used CEM to investigate the progressive collapse resistance of a thirty-
storey RC building under several column removal scenarios. The same researchers
developed a simplified method based on a two-scale model (Xue and Le, 2016a), which
was experimentally validated and applied to a ten-storey building (Xue and Le, 2016b).
Finally, in a context of design for robustness, several simplified analytical methods capable
of predicting the ultimate load and the collapse mode following column removal were
proposed in the literature. Due to their simplicity, analytical methods are particularly
suitable for conducting parametric analyses. In the case of multi-story RC frame with RC
flat slabs, the analytical method proposed by Martinelli et al. (2022) provides the ultimate
bearing capacity and the failure modes of a large variety of flat slabs. The analytical method
provides, in the form of design nomographs, the maximum load in accidental design
situation.
It should be emphasized that in relation to the application of numerical modelling strategies
further investigations are required in relation to the use of partial factors, the selection of
characteristic and representative values and load combinations.

5.2.3 Response of single components to abnormal loads


Structural robustness and progressive collapse should be investigated at the system level.
However, to this purpose, the response of single structural components to abnormal loads
has to be carefully modelled and analysed. This consideration applies to both key element
design and threat-dependent ALP analysis, where critical structural members (i.e. those
components the failure of which may activate progressive collapse) are analysed under
either specified or notional actions. Therefore, as outlined in several studies (e.g. Parisi
and Augenti, 2012), a local response analysis is carried out using either a linear or
nonlinear, static or dynamic analysis. Particularly under impulsive loading (associated with,
e.g., blast or impact), each single structural member can be assumed to be doubly-fixed
at its ends and can be ideally taken out from the entire structure because the latter does
not have sufficient time to develop global vibration modes, damping and inertia forces. In
line of principle, this assumption is met if the duration of loading is significantly lower than
the fundamental vibration period of the structure. Strain-rate effects can be explicitly
considered in dynamic analysis, provided that the structural software allows the user to
model strain-rate-dependent constitutive behaviour of materials. Otherwise, strain-rate
effects (particularly sensitivity of material strengths) can be indirectly incorporated in the

90
response analysis (be it static or dynamic) through dynamic increase factors (e.g. CEB,
1988; Malvar, 1999; Malvar and Ross, 1999). Even though various studies considered
strain-rate effects in subsequent response analysis of the residual structure, i.e. after
damage/removal of single/few members (e.g. Jayasooriya et al., 2012), they do not usually
influence the robustness of the structure.
If linear static analysis of a single structural member is carried out, strength demand in
terms of bending moment and shear force is simply compared to capacity. In this regard,
shear-flexure interaction should be taken in due consideration particularly in existing
structures not designed for earthquake resistance. Using nonlinear incremental static
analysis, the structural member is assumed to reach collapse if a plastic mechanism occurs
under the specified/notional load.
Linear or nonlinear dynamic analysis can be performed on one of the following capacity
models of the structural component (e.g. column, beam, slab, wall):
(1) a continuum model developed according to FEM (most used method) or other
numerical strategies presented in Sect. 5.2.2, or

(2) a single-degree-of-freedom (SDOF) system, which is a model recognised in guidelines


(e.g. ASCE, 1997, 2008; UFC 3-340-02, 2008) and widely investigated in several
studies (Nassr et al. 2012; Urgessa and Maji 2010; Wang et al. 2012).

In case (1), a doubly-fixed beam element with constant bending stiffness, distributed
mass, and zero or very low damping is widely used (e.g. Parisi and Augenti, 2012). Such
analysis then provides the motion of the dynamic system in terms of displacement,
velocity, and acceleration time histories, as well as strength and deformation demands to
be compared with capacity. In case of linear dynamic analysis, closed-form solutions are
readily available in the literature to predict the effects of different input time histories.
Thus, even linear dynamic analysis of continuum models can be sometimes sufficient to
assess the performance of single structural components, avoiding complex modelling
procedures and computations. This may be the case of, e.g., blast loading on building
columns, which can be basically analysed under triangular impulse uniformly distributed
over the column height, with peak overpressure and positive duration predicted through
simplified analytical models available in the literature (Parisi and Augenti, 2012).
Otherwise, nonlinear dynamic analysis of SDOF systems is a computationally efficient tool
(e.g. Dragos and Wu, 2013, 2015), but it may be affected by some limitations such as the
accurate prediction of shear and mixed failure modes of RC columns (e.g. Shi et al., 2008).
Based on such considerations, several researchers proposed pressure–impulse diagrams
to assess performance and damage of single structural members under different
assumptions regarding dynamic loading and structural behaviour (e.g. Fallah and Louca,
2007; Krauthammer et al., 2008; Shi et al., 2008; Park and Krauthammer, 2011; Ding et
al., 2013). The impulse is defined as the integral of pressure over the duration of the
pressure-time history. The pressure–impulse diagram is an iso-damage capacity curve of
the structural member defined by the combinations of pressure and impulse that produce
the same level of damage. Structural damage can be quantified through a scalar or vector-
valued measure, such as the percentage loss of axial load-bearing capacity of a column
(e.g. 20%, 50%, 80%). This highlights the need to define performance limit states
corresponding to increasing levels of structural damage. Therefore, the pressure–impulse
diagram is the boundary line of an iso-damage safety domain, meaning that dynamic loads
with pressure-impulse combinations falling below the pressure–impulse diagram do not
cause failure, and hence the prescribed level of damage considered in the performance
assessment. Pressure–impulse diagrams allow all types of failures to be considered, so
they are widely and easily used in research and engineering practice. Latest developments
have also produced probabilistic pressure–impulse diagrams for their use in performance-
based design and assessment of structural members (e.g. Parisi, 2015), as well as in
quantitative risk analysis of structures for robustness quantification.

91
Dealing with EC8-conforming RC frame structures subjected to blast scenarios, Parisi and
Augenti (2012) found that static, dynamic, and pressure–impulse analyses produce the
same safe/failed tagging of columns, delineating the same local damage scenarios for
subsequent global analysis of the structure. By contrast, static analysis produced too
conservative outcomes for buildings designed only to gravity loads, hence significantly
overestimating the number of failed elements under the same loading conditions.
The following sub-sections deal with the response of structural systems, which can be
analysed under either specified abnormal loads or notional scenarios. This is because
robustness assessment is not necessarily related to abnormal loads and structural collapse,
allowing the use of Level I methods in case of robustness quantification (see Sect. 5.2.4).
More sophisticated methods of structural response analysis can be applied to progressive
collapse analysis of structures, as described in Sects. 5.2.5 and 5.2.6.
It should be noted that in case of key element design, pushdown type analysis will not
necessarily highlight failures where load reversal occurs, e.g. in case of gas explosions.
Table 5-1 outlines the main advantages and drawbacks of the different methods, which
will be described in detail in the next sub-sections. That table particularly focuses on the
ease/complexity of structural modelling and computational cost, associated with
lacking/implicit/explicit consideration of material nonlinearities, second-order effects, and
dynamic effects (e.g. load amplification, inertia forces, damping forces).

Table 5-1. Some advantages and drawbacks of different methods of structural response analysis

Method Advantages Drawbacks


Linear static analysis Ease of structural modelling Implicit consideration of dynamic
Lowest computational cost effects through a dynamic
amplification factor
Simple safety checking
Lacking consideration of both
material and geometric
nonlinearities
Performance limit states to be
properly defined to ensure
representativeness for robustness
quantification
Linear dynamic Consideration of dynamic Lacking consideration of both
analysis effects (including dynamic material and geometric
load amplification, inertia nonlinearities
forces, and damping forces) Potential incorrect evaluation of
Moderate complexity of time- dynamic effects in case of structures
history analysis with large inelastic deformations
Moderate simplicity of safety Higher computational cost for
checking large/complex structures
Performance limit states to be
properly defined to ensure
conservative safety checks
Limit analysis Ease of structural modelling Lacking consideration of second-
Moderate computational cost order effects, to be included through
specific procedures
Simplified modelling of
material nonlinearities Threshold load factor to be properly
defined considering dynamic loading
Effective prediction of collapse
conditions
mechanisms
Nonlinear static Moderate ease of structural Inelastic capacity of structural
analysis modelling members to be properly defined
Moderate computational cost

92
Consideration of material and Dynamic amplification factor to
geometric nonlinearities define based on the expected level
Satisfactory prediction of of inelastic deformations
collapse mechanisms
Nonlinear dynamic Consideration of material and Sophistication of structural
analysis geometric nonlinearities modelling (including damping and
Explicit evaluation of dynamic inertia masses)
effects High computational cost
Realistic prediction of
nonlinear dynamic response
and collapse mechanisms
Probabilistic analysis Explicit consideration of Highest computational cost
uncertainties
Rational assessment of
progressive collapse
resistance and robustness
Possible use of analysis output
in decision-making for disaster
risk mitigation

5.2.4 Linear and limit analysis methods for structural systems (Level I)
The structural response to different hazardous scenarios can be investigated through either
linear or limit analysis methods, which are computationally efficient and easy to use in
engineering practice. These methods can effectively be used for robustness quantification.
Indeed, response of single components plays a role when assessing progressive collapse
against abnormal loads, but it is not required in other hazardous scenarios.
Linear analysis procedures are based upon the assumption of linear elastic behaviour of
materials, provided that their mechanical strength is properly scaled down to a design
value via partial safety factors that account for both material and model uncertainties.
Structural response can be evaluated either statically or dynamically, in the former case
using load amplification factors to define equivalent static loads. This approach can
extremely useful to compare the structural response to different damage scenarios and,
therefore, to quantify structural robustness (Powell 2009).
Linear static analysis (LSA) procedures require the assumption of dynamic amplification
factors for gravity loads, the definition of which can be a critical issue depending on the
expected level of ductility demand on structural components and the availability of specific
formulations for the structure under consideration. In robustness quantification and
progressive collapse analysis, dynamic amplification factors are applied to gravity loads in
structural components (e.g. frame members, floor systems) that are directly involved in
the dynamic response of the structure to damage. In the frequent case of buildings for
which robustness to notional removal of single or few structural components (e.g. columns,
walls) has to be evaluated, dynamic amplification factors are applied to increase the
intensity of gravity loads in floor areas above the removed components. This is because
inertia masses in those floor areas are accelerated by the sudden failure of those structural
elements, and the amount of vertical acceleration those masses are subjected to depends
on the level of inelasticity that develops within the structure (particularly in beams and
floor systems). Accordingly, Brunesi and Parisi (2017) derived a set of regression models
for both gravity-load designed and earthquake-resistance RC frame buildings designed
according to Eurocodes, because no formulation is still available in those codes for
European structures. The same applies to other structural types, as pointed out by
Mucedero et al. (2021).

93
As pointed out by Marjanishvili (2004), LSA is the simplest method of structural response
analysis and it usually leads to conservative predictions of strength demand on structural
elements. The advantages of LSA include relative simplicity of structural modelling, quick
calculations for structural response analysis, and ease of safety checking, the latter
involving force-based verifications of structural components such as beams in frame
structures. By contrast, LSA has several disadvantages that include the lacking
consideration of both material and geometric nonlinearities, as well as dynamic effects
such as inertia forces and damping. Therefore, LSA can be a very useful tool for simple
structures with predictable behaviour, while calling for careful use on complex and large
structures that should be evaluated with nonlinear analysis methods.
Linear dynamic analysis (LDA) allows consideration of dynamic effects, while material and
geometric nonlinearity effects can be indirectly considered. LDA is typically carried out
when assessing structural robustness under sudden loss of major load-bearing elements,
which produces dynamic motion of the structural system. More specifically, a time-history
analysis of the structure is performed, for instance using direct integration methods for
solving the equations of motion. LDA is deemed more accurate than its LSA counterpart
because it accounts for dynamic amplification, inertia forces, and damping forces, while
ensuring a moderate complexity of calculations. In addition to the lacking consideration of
nonlinearity effects in the modelling and analysis of the structure, LDA has the
disadvantage of being more time consuming for large structures. It is also noted that
dynamic amplification and both inertia and damping forces may be incorrectly evaluated
for structures that develop large inelastic deformations. Force-based safety checks are
performed, as observed in the case of LSA. Maximum strength demands (in terms of
internal forces such as bending moments and shear forces in frame members) are
evaluated over the entire duration of the structural analysis. Performance evaluation
criteria can be assumed to be rather conservative for structures with nearly elastic
behaviour, while they could become non-conservative in the case of structures that are
expected to experience large inelastic deformations. Marjanishvili and Agnew (2006)
showed that dynamic analysis procedures not only yield more accurate response
predictions but allow ease of use in progressive collapse assessment. Nonetheless,
performance limit states for linear analysis should be properly defined to avoid
unconservative safety checks and robustness quantification. Otherwise, a structure that
meets performance criteria in linear analysis may exceed performance limits in nonlinear
dynamic analysis.
In addition to linear elastic analysis methods, the structural behaviour can be evaluated by
means of limit analysis (LA) that makes use of rigid-plastic models. This modelling strategy
is relatively simple to use in several software packages, even allowing hand calculations
for relatively simple structures. LEA can consist of plastic analysis (e.g.: Watwood, 1979;
Corotis and Nafday, 1990), which searches for the load factor on the applied loads for
which the following requirements are met: equilibrium equations are satisfied (static
condition), and a sufficient number of plastic hinges are formed in the structure to activate
a partial or total collapse mechanism (kinematic condition). For instance, in the case of a
frame structure, nonlinear behaviour of the structure is lumped at the ends of each frame
member, which is assumed to develop its ultimate bending moment accounting for its
interaction with axial and shear forces. It could be shown that LA turns out to be a linear
optimization programming problem, the objective of which is to minimise the load factor
(e.g. Grierson and Gladwell, 1971). This optimization problem can be solved through a
simplex algorithm, assuming independent mechanisms. In the case of a frame structure,
these collapse mechanisms can be soft-storey mechanisms, beam mechanisms, or joint
mechanisms. Whilst in static loading conditions the structure is assumed to collapse when
the load factor is less than or equal to 1, in dynamic loading conditions the threshold load
factor can be conservatively set to 2. Ruth et al. (2006) found that the actual load factor
associated with structural instability under gravity loads is between 1 and 2. Limit analysis
can easily incorporate the effects of aging and deterioration processes and be applied to
assess the time-variant load capacity of deteriorating systems (Biondini & Frangopol 2008).
Second-order internal forces that can prevent the activation of a collapse mechanism can

94
be taken into account in LA as shown e.g. by Park and Gamble (2000) for arching in axially
restrained reinforced concrete elements and by Sawczuk and Winnicki (1965) or Herraiz
and Vogel (2016) for tensile membrane action in reinforced concrete slabs.

5.2.5 Nonlinear analysis methods for structural systems (Level II)


If a progressive collapse analysis of the structure has to be performed, a nonlinear analysis
method is required. In such a context, the capacity model of the structural system should
be based on realistic assumptions regarding, e.g., rotational capacity of plastic hinges or
ultimate strains of individual materials depending on whether a lumped or spread plasticity
approach is used. This motivates the upper-bound values prescribed by some building
codes and guidelines for rotational capacity of plastic hinges in different types of structures
(e.g. steel frames, concrete frames). It is evident that too large capacity values may result
in unrealistic, non-conservative design/assessment solutions for progressive collapse
resistance, such as: (i) small tying systems associated with too large rotational capacity
values assigned to beams (in indirect design methods); and (ii) very low values of dynamic
amplification factors for gravity loads on floor areas above locally damaged/removed
elements (in direct design methods based on static response analysis). Another remark
involves the role of strain rate effects, which do not typically play a key role when assessing
the progressive collapse resistance of the structure, as opposed to the case of single
components subjected to impulsive loading (e.g. impact, bomb detonation). Lastly, it
should be emphasized that not all detailed material properties are specified in commonly
available product specifications.
Nonlinear incremental static (pushdown) analysis, such as the one proposed by Izzuddin
et al. (2008) in their ductility-based robustness assessment methodology, was recognised
to be a powerful tool and was applied to several types of structures (e.g. Vlassis et al.,
2008; Khandelwal and E-Tawil, 2011; Mucedero et al., 2021). De Biagi et al. (2017)
developed a pushdown-based method where structural robustness can be evaluated under
progressive damage to structural members of RC structures. This allows effects of
degradation to be incorporated in robustness assessment. Based on pushdown analysis,
the robustness of the structure subjected to local damage can be quantified starting from
the ultimate load factor of gravity loads. It is recalled that static analysis procedures require
the assumption of dynamic amplification factors for gravity loads, as noted in Sect. 5.2.4.
Alternatively, in order to avoid the use of dynamic amplification factors or dynamic
analyses, the energy-based method was developed, which allows to derive a dynamic
capacity curve from a static (pushdown) analysis (Xu and Ellingwood, 2011).
As discussed by, e.g., Arup (2011) and Byfield et al. (2014), nonlinear time history analysis
has the highest level of complexity, but it explicitly visualises all features of the structural
behaviour. Implicit solution algorithms in nonlinear time history analysis are generally
affected by convergence issues that can be usually overcome by explicit algorithms,
resulting in more robust and accurate computations. Brunesi et al. (2015) proposed and
implemented incremental dynamic analysis (IDA) for robustness assessment of structures,
where inertia masses associated with gravity loads are gradually increased to assess
whether progressive collapse occurs or not. To that aim and based on IDA, Parisi et al.
(2019) defined a set of performance limit states for progressive collapse analysis, which
could be considered in ALP methods for robustness assessment of RC structures.
Depending on the performance level under consideration (from slight damage to near
collapse), structure response in the range of large deformations is sensitive to different
properties (e.g. concrete compressive strength, steel yield strength, beam span length,
ultimate/fracture steel strain). This highlights that the accuracy level of capacity modelling
should be defined on the basis of the target performance to assess.

5.2.6 Probabilistic analysis methods for structural systems (Level III)


The use of different nonlinear analysis methods, design provisions (e.g. seismic or non-
seismic design criteria and detailing), and capacity models (e.g. 2D or 3D) can produce
different predictions on the same structure. This applies to the response analysis of both

95
structural components and entire structures, resulting in different quantifications of
element resistance and system robustness.
The structure can be analysed accounting for different sources of uncertainty, such as
those involving material properties, geometric parameters, loads, capacity modelling,
exposure scenario and deterioration processes. In that context, the progressive collapse
resistance and robustness of a structure can be quantified through simulation procedures
such as the Monte Carlo method. This allows the structural engineer to assess the
probability of failure, which should not be greater than a target probability. In threat-
independent scenarios (see e.g. Parisi, 2015), the collapse fragility of the structure can be
decomposed in the fragility of single components (i.e. the conditional probability of
component failure given hazard) and the progressive collapse fragility of the structural
system (i.e. the conditional probability of system failure given component damage).
Comparing the failure probability of the structural system to that of single components
allows robustness quantification through the computation of probability-based measures
(see Sect. 5.1.2). Furthermore, (conditional) risk-based measures can be efficiently
calculated using efficient simulation procedures such as Latin Hypercube sampling, as e.g.
shown in (Droogné et al., 2018) where the structure is decomposed in two parts to further
reduce the computational cost of such analyses.
Hence, uncertainties affect structural response at both component and system scales.
Parisi et al. (2015) developed fragility surfaces of RC columns subjected to blast loading,
using a pressure–impulse formulation proposed by Shi et al. (2008). Each fragility surface
provides the conditional probability of exceeding a prescribed level of damage (e.g. 20%
loss of axial load-bearing capacity) given the peak overpressure and impulse of blast load.
Uncertainty in material strengths, column dimensions, reinforcement ratios, and blast
capacity model were modelled and propagated. Horizontal sections of fragility surfaces at
different probability levels produced performance-based pressure–impulse diagrams for
their use in research and engineering practice. Both fragility surfaces and pressure–impulse
diagrams demonstrated that RC columns of earthquake-resistant buildings have high
resistance to blast loading. Fragility surfaces of RC columns were more recently used to
evaluate the risk-targeted safety distance of RC frame buildings from natural-gas pipelines
(Russo and Parisi, 2016) and assess potential damage due to hydrogen pipeline explosions
(Russo et al., 2019).
Brunesi et al. (2015) derived fragility functions for European, low-rise, RC framed buildings
representative of both gravity-load and seismically designed structures according to
Eurocodes. Fragility functions corresponding to multiple damage states indicated a
significant influence of seismic design/detailing on robustness of RC buildings. Structural
robustness was thus probabilistically evaluated through fragility functions, which provide
the conditional probability of exceeding a damage level given the magnitude of gravity
loads. The inclusion of secondary beams in the capacity model of the structure also played
a significant role, notably increasing the building robustness because of the increased
capability of redistributing loads from areas directly involved in progressive collapse
mechanisms.
The influence of uncertainties in robustness quantification has been investigated also in
Biondini and Frangopol (2014), where a simulation-based probabilistic analysis is applied
to time-variant robustness of concrete structures under reinforcement corrosion.
Botte et al. (2021) determined the sensitivity of the development of arching or catenary
actions in RC elements and showed that variables which are usually not explicitly taken
into account by traditional design methods, such as the axial restraint stiffness and ultimate
reinforcement strain have a significant influence on the resistance of such elements.
Such studies further remarked that the accuracy level of capacity modelling and response
analysis methods has significant impact on safety assessment and robustness
quantification. It is also emphasised that effects of degradation may also notably affect the
output of fragility analysis, but this point needs to be investigated to provide detailed
information and to draw some conclusions.

96
5.3 Damage-based robustness quantification
Several robustness indicators have been presented that express the variation of system
performance induced by initial damage compared with the intact system in terms of risk,
vulnerability, redundancy, residual strength, and damage tolerance. These robustness
indicators can effectively be used to compare in a relative way the system robustness with
respect to different damage scenarios and exposures. However, it is worth noting that such
indicators are, in general, not sufficient to enable comparisons between different structural
systems or ranking of different design alternatives. To this purpose, the residual
performance indicator 0≤≤1 associated with the damage scenario should be related with
the amount of damage 0≤≤1 to provide a functional  (). Performance indicators 𝜌
can be referred to serviceability or ultimate limit states and the damage index  should be
computed taking into account the spatial distribution of damage and the role of different
component materials in non-homogeneous systems, as already discussed. Depending on
the purpose of the robustness analysis, this procedure can be applied using a deterministic
approach or taking uncertainties and consequences into account with probability-based,
reliability-based, and risk-based approaches.
The following criterion has been proposed in Biondini (2009) to quantify structural
robustness:

𝑅(𝜌, 𝛥) = 𝜌𝛼 + Δ𝛼 ≥ 1 (5-22)

where 𝑅 = 𝑅(𝜌, Δ) is a robustness factor, and 𝛼 is a shape parameter of the boundary


𝑅(𝜌, Δ) = 1. The structural system is robust when the criterion is satisfied (𝑅 > 1), and not
robust otherwise (𝑅 < 1). This concept is illustrated in Figure 5-4.a for the case 𝛼 = 1. As
shown in Figure 5-4.b, the value of the parameter 𝛼 can be properly selected according to
the acceptable level of damage susceptibility for the structure under investigation. A value
𝛼 = 1, which indicates a proportionality between acceptable loss of performance and
damage, should be appropriate in most cases. Values 𝛼 < 1 should be avoided since allow
for disproportionate damage effects. On the other hand, values 𝛼 > 1 can be required for
structures of strategic importance.

Figure 5-4. Robustness factor 𝑅 = 𝑅(𝜌, Δ). (a) Performance 𝜌 vs damage Δ state diagram (𝛼 = 1);
(b) Role of the shape parameter 𝛼 on the robustness threshold 𝑅 = 1.
(Biondini and Frangopol, 2014)

(a) (b)

A proper value of the parameter 𝛼 can be selected based on a threshold value of the area
𝐴 = 𝐴(𝛼) [0,1] lying under the curve 𝑅 = 1 (Di Silvestri et al. 2014):

97
1 1 1
𝐴(𝛼) = {∫ 𝜌(𝛼, Δ)𝑑Δ | 𝑅 = 1} = ∫ (1 − Δ𝛼 )𝛼 𝑑Δ (5-23)
0 0

which leads to 𝐴 = 0 for 𝛼 = 0, 𝐴 = 0.5 for 𝛼 = 1, and 𝐴 = 1 for 𝛼 = .


The importance factor 𝛼 emphasizes that the robustness measure should depend not only
on system properties and damage mechanisms, but also on the importance of the system.
This approach can be effectively used to compare the robustness associated to different
structural systems and rank design alternatives. Moreover, it can accommodate damage
scenarios with sudden damage or continuous damage and life-cycle formulations based on
time-variant state variables, i.e. time-variant damage index Δ = Δ(𝑡), performance index
𝜌 = 𝜌(𝑡), and robustness factor 𝑅 = 𝑅(𝑡) (Biondini 2009, Biondini & Frangopol 2014).
An illustrative example is qualitatively presented in Figure 5-5 for two different damage
scenarios in terms of functions  () (Figure 5-5(a)) and time-variant robustness 𝑅 =
𝑅(𝑡) for =1 (Figure 5-5(b)). In this example, scenario #2 with lower damage rate is more
detrimental to robustness compared to scenario #1. Using this approach, the robustness
factor may also provide useful information to plan repair interventions and maintenance
actions to ensure long-term robustness.

Figure 5-5. Comparison of time-variant robustness over a structural lifetime T for two damage
scenarios (indicators at equal time intervals t=T/5): (a) performance functions  (), and
(b) robustness factor profiles 𝑅 = 𝑅(𝑡) for 𝛼 = 1.

(a) (b)
Damage-based integral measures of structural robustness, averaging the robustness factor
over damage extension or time intervals, are also proposed in literature (Starossek &
Haberland 2011, Cavaco et al. 2018). However, it is worth noting that the relationship
R=R(,) is time-variant and nonlinear and the robustness criterion R(t)1 needs to be
verified at discrete points in time over the structural life-cycle. In fact, integral measures
of robustness based on the following formulation:

R    ()d
1
(5-24)
0

should be avoided since they provide only average indications over the lifetime and are not
able to describe the actual level of structural robustness. Examples can be found in
Starossek & Haberland (2011).

98
References
André, J., Beale, R., Baptista, A., ‘New indices of structural robustness and structural
fragility’, Structural Engineering and Mechanics, Vol. 56, No 6, 2015, pp. 1063–1093.
Adam, J., Parisi, F., Sagaseta, J., Lu, X. ‘Research and practice on progressive collapse and
robustness of building structures in the 21st century’, Engineering Structures, 173, 122-
149, 2018.
Alashker, Y., Li, H., El-Tawil, S. ‘Approximations in progressive collapse modelling’, Journal
of Structural Engineering, 137(9), 914-924, 2011.
Arup. ‘Review of International Research on Structural Robustness and Disproportionate
Collapse’, Department for Communities and Local Government, London, 2011.
ASCE ‘Design of blast resistant buildings in petrochemical facilities’, American Society of
Civil Engineers, Reston, VA, 1997.
ASCE/SEI 59-11 ‘Blast protection of buildings’, American Society of Civil Engineers, Reston,
VA, 2008.
Baker, J.W., Schubert, M., Faber, M.H., ‘On the assessment of robustness’, Journal of
Structural Safety, vol. 30, pp. 253-267, 2008.
Bao, Y., Kunnath, S.K., El-Tawil, S., Lew, H.S. ‘Macromodel-based simulation of
progressive collapse: RC frame structures’, Journal of Structural Engineering, 134(7),
1079-1091, 2008.
Biondini, F., ‘A Measure of Lifetime Structural Robustness’, ASCE/SEI Structures Congress
2009, Austin, TX, USA, April 30-May 2, 2009. In: Proceedings of the Structures Congress
2009, L. Griffis, T. Helwig, M. Waggoner, M. Hoit (Eds.), American Society of Civil
Engineers, USA, 2009.
Biondini, F., Discussion: ‘Time-variant redundancy of ship structures’, by Decò, A.,
Frangopol, D.M., & Okasha, N.M., Society of Naval Architects and Marine Engineers
Transactions, SNAME, Volume 119, page 40, 2012.
Biondini, F., Camnasio, E., Palermo, A. ‘Lifetime seismic performance of concrete bridges
exposed to corrosion’, Structure and Infrastructure Engineering, Taylor & Francis, 10(7),
2014, 880-900.
Biondini, F., Frangopol, D. M. ‘Probabilistic limit analysis and lifetime prediction of concrete
structures’. Structure and Infrastructure Engineering, 4(5), 399-412, 2008.
Biondini, F., Frangopol, D.M., ‘Time-variant Robustness of Ageing Structures’, Chapter
6, Maintenance and Safety of Ageing Infrastructure, Y. Tsompanakis & D.M. Frangopol
(Eds.), Structures and Infrastructures Book Series, CRC Press, Taylor & Francis Group, 10,
163-200, 2014.
Biondini, F., Frangopol, D.M., ‘Time-variant Redundancy and Failure Times of Deteriorating
Concrete Structures considering Multiple Limit States’, Structure and Infrastructure
Engineering, Taylor & Francis, 13(1), 94-106, 2017.
Biondini, F., Frangopol, D.M., Restelli, S., ‘On Structural Robustness, Redundancy and
Static Indeterminacy’, ASCE Structures Congress 2008, Vancouver, B.C., Canada, 2008.
Biondini, F., Restelli, S., ‘Damage Propagation and Structural Robustness’, First
International Symposium on Life-Cycle Structural Engineering (IALCCE'08), Varenna, Italy,
June 10-14, 2008. In Life-Cycle Civil Engineering, F. Biondini and D.M. Frangopol (Eds.),
CRC Press, Taylor and Francis Group, 131-136.
Biondini, F., Vergani, M., “Deteriorating beam finite element for nonlinear analysis of
concrete structures under corrosion” Structure and Infrastructure Engineering, 11(4),
2015, 519-532.

99
Botte, W., Droogné, D., Caspeele R. ‘Reliability-based resistance of RC element subjected
to membrane action and their sensitivity to uncertainties’, Engineering Structures, 238,
112259, 2021.
Brunesi, E., Nascimbene, R. ‘Extreme response of reinforced concrete buildings through
fiber force-based finite element analysis’, Engineering Structures, 69, 206-215, 2014.
Brunesi, E., Nascimbene, R., Parisi, F., Augenti, N. ‘Progressive collapse fragility of
reinforced concrete framed structures through incremental dynamic analysis’, Engineering
Structures, 104, 65-79, 2015.
Brunesi, E., Parisi, F., ‘Progressive collapse fragility models of European reinforced concrete
framed buildings based on pushdown analysis’, Engineering Structures, 152, 579-596,
2017.
Byfield, M., Mudalige, W., Morison, C., Stoddart, E. ‘A review of progressive collapse
research and regulations’, Proceedings of the Institution of Civil Engineers – Structures and
Buildings, 167(SB8), 447-456, 2014.
Cavaco, E. S., Neves, L. A., Casas, J. R., ‘On the robustness to corrosion in the life cycle
assessment of an existing reinforced concrete bridge’, Structure and Infrastructure
Engineering, 14(2), 137-150, 2018.
CEB Bulletin 187 ‘Concrete structures under impact and impulsive loading’, Comité Euro-
International du Béton, Lausanne, 1988.
Corotis R.B., Nafday A.M. ‘Application of mathematical programming to system reliability’,
Structural Safety, 7, 149-154, 1990.
De Biagi, V., Parisi, F., Asprone, D., Chiaia, B., Manfredi, G. ‘Collapse resistance
assessment through the implementation of progressive damage in finite element codes’,
Engineering Structures, 136, 523-534, 2017.
Di Silvestri, V., Biondini, F., Titi, A., Frangopol, D. M., ‘Lifetime robustness of a RC bridge
pier under corrosion considering bridge importance’, Seventh International Conference on
Bridge Maintenance, Safety, and Management (IABMAS 2014), Shanghai, China, July 7-11,
2014. In: Bridge Maintenance, Safety, Management and Life Extension, A. Chen, D.M.
Frangopol, X.Ruan, (Eds.), CRC Press/Balkema, Taylor & Francis Group, 1699-1704, 2014.
Ding, Y., Wang, M., Li, Z.-X., Hao H. ‘Damage evaluation of the steel tubular column
subjected to explosion and post-explosion fire condition’, Engineering Structures, 55, 44–
55, 2013.
Dinu, F., Marginean, I., Dubina, D., Petran, I. ‘Experimental testing and numerical analysis
of 3D steel frame system under column loss’, Engineering Structures, 113, 59-70, 2016.
Dragos, J., Wu, C. ‘A new general approach to derive normalised pressure impulse curves’,
International Journal of Impact Engineering, 62, 1-12, 2013.
Dragos, J., Wu, C. ‘Single-degree-of-freedom approach to incorporate axial load effects on
pressure impulse curves for steel columns’, Journal of Engineering Mechanics, 141(1),
04014098, 2015.
Droogné, D., Botte, W., Caspeele R. ‘A multilevel calculation scheme for risk-based
robustness quantification of reinforced concrete frames’, Engineering Structures, 160, 56-
70, 2018.
Elsanadedy, H.M., Almusallam, T.H., Al-Salloum, Y.A., Abbas, H. ‘Investigation of precast
RC beam-column assemblies under column loss scenario’, Construction and Building
Materials, 142, 552-571, 2017.
El-Tawil, S., Li, H., Kunnath, S. ‘Computational simulation of gravity-induced progressive
collapse of steel-frame buildings: Current trends and future needs’, Journal of Structural
Engineering, 140(8), 1-12, 2014.

100
Soleiman Fallah, A., Louca, L.A. ‘Pressure–impulse diagrams for elastic-plastic-hardening
and softening single-degree-of-freedom models subjected to blast loading’, International
Journal of Impact Engineering, 34, 823–842, 2007.
Feng, D.-C., Xiong, C.-Z., Brunesi, E., Parisi, F., Wu, G. ‘Numerical simulation and
parametric analysis of precast concrete beam-slab assembly based on layered shell
elements’, Buildings, 11(1), 7, 2021.
Frangopol, D.M., Curley, J.P., ‘Effects of Damage and Redundancy on Structural Reliability’,
ASCE Journal of Structural Engineering, 113(7), 1533-1549, 1987.
Fu, F. ‘3-D nonlinear dynamic progressive collapse analysis of multi-storey steel composite
frame buildings – Parametric study’, Engineering Structures, 32(12), 3974-3980, 2011.
Ghosn, M., Moses, F., ‘NCHRP Report 406: Redundancy in highway bridge superstructures’.
Washington, DC: Transportation Research Board, National Research Council, 1998.
Grierson D.E., Gladwell G.M.L. ‘Collapse load analysis using linear programming’, Journal
of Structural Division, 97(ST5), 1561-1573, 1971.
Gu, X., Wang, X., Yin, X., Lin, F., Hou, J. ‘Collapse simulation of reinforced concrete
moment frames considering impact actions among blocks’, Engineering Structures, 65, 30-
41, 2014.
Guo, L., Gao, S., Fu, F., Wang, Y. ‘Experimental study and numerical analysis of
progressive collapse resistance of composite frames’, Journal of Constructional Steel
Research, 89, 236-251, 2013.
Izzuddin, B.A., Vlassis, A.G., Elghazouli, A.Y., Nethercot, D.A. ‘Progressive collapse of
multi-storey buildings due to sudden column loss – Part I: Simplified assessment
framework’, Engineering Structures, 30(5), 1308-1318, 2008.
Herraiz, B., Vogel, T. ‘Novel design approach for the analysis of laterally unrestrained
reinforced concrete slabs considering membrane action’ Engineering Structures, 123, 313-
329, 2016.
Jahromi, H.Z., Vlassis, A.G., Izzuddin, B.A. ‘Modelling approaches for robustness
assessment of multi-storey steel-composite buildings’, Engineering Structures, 51, 278-
294, 2013.
Jayasooriya, R., Thambiratnam, D.P., Perera, N.j., Kosse, V. ‘Blast and residual capacity
analysis of reinforced concrete framed buildings’, Engineering Structures, 33(12), 3483–
3495, 2011.
Kazemi-Moghaddam, A., Sasani, M. ‘Progressive collapse evaluation of Murrah Federal
Building following sudden loss of column G20’, Engineering Structures, 89, 162-171, 2015.
Keyvani, L., Sasani, M. ‘Analytical and experimental evaluation of progressive collapse
resistance of a flat-slab posttensioned parking garage’, Journal of Structural Engineering,
141(11), 1-8, 2015.
Khalil, A.A. ‘Enhanced modeling of steel structures for progressive collapse analysis using
the Applied Element Method’, Journal of Performance of Constructed Facilities, 26(6), 766-
779, 2012.
Khandelwal, K., E-Tawil, S. ‘Pushdown resistance as a measure of robustness in
progressive collapse’, Engineering Structures, 33(9), 2653-2661, 2011.
Krauthammer, T., Astarlioglu, S., Blasko, J., Soh, T.B., Ng, P.H. ‘Pressure–impulse
diagrams for the behavior assessment of structural components’, International Journal of
Impact Engineering, 35, 771-783, 2008.
Kunnath, S.K., Bao, Y., El-Tawil, S. ‘Advances in computational simulation of gravity-
induced disproportionate collapse of RC frame buildings’, Journal of Structural Engineering,
144(2), 1-18, 2018.

101
Le, J.L., Xue, B. ‘Probabilistic analysis of reinforced concrete frame structures against
progressive collapse’, Engineering Structures, 76, 313-323, 2014.
Li, S., Shan, S., Zhai, C., Xie, L. ‘Experimental and numerical study on progressive collapse
process of RC frames with full-height infill walls’, Engineering Failure Analysis, 59, 57-68,
2016.
Liu, Y., Xu, L., Grierson, D.E. ‘Influence of semi-rigid connections and local joint damage
on progressive collapse of steel frameworks’, Computer-Aided Civil and Infrastructure
Engineering, 25(3), 184-204, 2010.
Lu, X., Li, Y., Guan, H., Ying, M. ‘Progressive collapse analysis of a typical super-tall
reinforced concrete frame-core tube building exposed to extreme fires’, Fire Technology,
53(1), 107-133, 2017.
Lu, X., Lin, X., Ye, L. ‘Simulation of structural collapse with coupled finite element-discrete
element method’, Computational Structural Engineering, Springer Netherlands, 127-135,
2009.
Maes, M. A., Fritzsons, K. E., Glowienka, S., ‘Structural robustness in the light of risk and
consequence analysis’, Structural Engineering International, 16(2), 101-107, 2006.
Malvar, L.J. ‘Review of static and dynamic properties of steel reinforcing bars’, ACI
Materials Journal, 96(5), 609–616, 1999.
Malvar, L.J., Ross, C.A. ‘Review of strain rate effects for concrete in tension’, ACI Materials
Journal, 96(5), 735–739, 1999.
Marjanishvili, S.M. ‘Progressive analysis procedure for progressive collapse’, Journal of
Performance of Constructed Facilities, 18(2), 79-85, 2004.
Marjanishvili, S., Agnew E. ‘Comparison of various procedures for progressive collapse
analysis’, Journal of Performance of Constructed Facilities, 20(4), 365-374, 2006.
Martinelli P., Colombo M., Ravasini S., Belletti B. ‘Application of an analytical method for
the design for robustness of RC flat slab buildings’, Engineering Structures, 258, 114117,
2022
Masoero, E., Wittel, F.K., Herrmann, H.J., Chiaia, B. ‘Progressive collapse mechanisms of
brittle and ductile framed structures’, Journal of Engineering Mechanics, 136(8), 987-995,
2010.
Mohammad El Hajj Diab, André Orcesi, Cédric Desprez, Jérémy Bleyer. A progressive
collapse modelling strategy coupling the yield design theory with non-linear analysis.
Engineering Structures, 2021, 16 p. ⟨10.1016/j.engstruct.2020.111832⟩.

Mohammad El Hajj Diab, Cédric Desprez, André Orcesi, Jeremy Bleyer. Structural
robustness quantification through the characterization of disproportionate collapse
compared to the initial local failure. Engineering Structures, Elsevier, 2022, 255,
pp.113869. ⟨10.1016/j.engstruct.2022.113869⟩.
Mucedero, G., Brunesi, E., Parisi, F. ‘Nonlinear material modelling for fibre-based
progressive collapse analysis of RC framed buildings’, Engineering Failure Analysis, 118,
104901, 2020.
Mucedero, G., Brunesi, E., Parisi, F. ‘Progressive collapse resistance of framed buildings
with partially encased composite beams’, Journal of Building Engineering,
https://doi.org/10.1016/j.jobe.2021.102228, 2021.
Nassr, A.A., Razaqpur, A.G., Tait, M.J., Campidelli, M., Foo, S. ‘Experimental performance
of steel beams under blast loading, Journal of Performance of Constructed Facilities, 26(5),
600-619, 2012.
Parisi, F. ‘Blast fragility and performance-based pressure–impulse diagrams of European
reinforced concrete columns’, Engineering Structures, 103, 285-297, 2015.

102
Parisi, F., Augenti, N. ‘Influence of seismic design criteria on blast resistance of RC framed
buildings: A case study’, Engineering Structures, 44, 78–93, 2012.
Parisi, F., Balestrieri, C., Asprone, D. ‘Blast resistance of tuff stone masonry walls’,
Engineering Structures, 113, 233-244, 2016.
Parisi, F., Scalvenzi, M. ‘Progressive collapse assessment of gravity-load designed
European RC buildings under multi-column loss scenarios’, Engineering Structures, 209,
110001, 2020.
Parisi, F., Scalvenzi, M., Brunesi, E. ‘Performance limit states for progressive collapse
analysis of reinforced concrete framed buildings’, Structural Concrete, 20(1), 68-84, 2019.
Park, J.Y, Krauthammer, T. ‘Pressure impulse diagrams for simply-supported steel columns
based on residual load-carrying capacities’, Structural Engineering and Mechanics, 39(2),
287-301, 2011.
Park, R., Gamble, W.L. ‘Reinforced Concrete Slabs’, John Wiley & Sons Inc., New York (US),
2000.
Pekau, O.A, Cui, Y. ‘Progressive collapse simulation of precast panel shear walls during
earthquakes’, Computers and Structures, 84, 400-412, 2006.
Pham, A.T., Lim, N.S., Tan, K.H. ‘Investigations of tensile membrane action in beam-slab
systems under progressive collapse subject to different loading configurations and
boundary conditions’, Engineering Structures, 150, 520-536, 2017.
Powell, G., ‘Disproportionate Collapse: The futility of using nonlinear analysis’, ASCE/SEI
Structures Congress 2009, Austin, TX, USA, April 30-May 2, 2009. In: Proceedings of the
Structures Congress 2009, L. Griffis, T. Helwig, M. Waggoner, M. Hoit (Eds.), American
Society of Civil Engineers, USA, 2009.
Ravasini, S., Belletti, B., Brunesi, E., Nascimbene, R., Parisi, F. ‘Nonlinear dynamic
response of a precast concrete building to sudden column removal’, Applied Sciences, 11,
599, 2021.
Russo, P., De Marco, A., Parisi, F. ‘Failure of reinforced concrete and tuff stone masonry
buildings as consequence of hydrogen pipeline explosions’, International Journal of
Hydrogen Energy, 44(38), 21067-21079, 2019.
Russo, P., Parisi, F. ‘Risk-targeted safety distance of reinforced concrete buildings from
natural-gas transmission pipelines’, Reliability Engineering and System Safety, 148, 57-
66, 2016.
Ruth P., Marchand K.A., Williamson E.B. ‘Static equivalency in progressive collapse
alternate path analysis: reducing conservatism while retaining structural integrity’, Journal
of Performance of Constructed Facilities, 20(4), 349-364, 2006.
Sadek, F., El-Tawil, S., Lew, H.S. ‘Robustness of composite floor systems with shear
connections: Modeling, simulation and evaluation’, Journal of Structural Engineering,
134(11), 1717-1725, 2008.
Sadek, F., Main, J.A., Lew, H.S., Bao, Y. ‘Testing and analysis of steel and concrete beam-
column assemblies under a column removal scenario’, Journal of Structural Engineering,
137(9), 881-892, 2011.
Salem, H., Mohssen, S., Nishikiori, Y., Hosoda, A. ‘Numerical collapse analysis of
Tsuyagawa bridge damaged by Tohoku tsunami’, Journal of Performance of Constructed
Facilities, 30(6), 1-12, 2016.
Salem, H.M., Helmy, H.M. ‘Numerical investigation of collapse of the Minnesota I-35W
bridge’, Engineering Structures, 59(2), 635-645, 2014.
Sawczuk, A., Winnicki, L. ‘Plastic behavior of simply supported reinforced concrete plates
at moderately large deflections’, International Journal of Solids and Structures 1, 97-111,
1965.

103
Shi, Y., Hao, H., Li, Z.-X. ‘Numerical derivation of pressure–impulse diagrams for prediction
of RC column damage to blast loads’, International Journal of Impact Engineering, 35(11),
1213–1227, 2008.
Shi, Y., Li, Z.X., Hao, H. ‘A new method for progressive collapse analysis of RC frames
under blast loading’, Engineering Structures, 32(6), 1691-1703, 2010.
Smith, J. W., ‘Structural robustness analysis and the fast fracture analogy’, Structural
Engineering International, 16(2), 118-123, 2006.
Starossek, U., Haberland, M., ‘Approaches to measure of structural robustness’, Structure
and Infrastructure Engineering, 625-631, 2011.
Starossek, U., Haberland, M., ‘Measures of structural robustness – Requirements &
applications’, ASCE Structures Congress 2008, Vancouver, B.C., Canada, 2008.
Sun, R., Burgess, I.W., Huang, Z., Dong, G. ‘Progressive failure modelling and ductility
demand of steel beam-to-column connections in fire’, Engineering Structures, 89(17), 66-
78, 2015.
Tagel-Din, H., Meguro, K. ‘Applied Element Method for simulation of nonlinear materials:
Theory and application for RC structures’, Structural Engineering/Earthquake Engineering,
Japan Society of Civil Engineers, 17(2), 137-148, 2000.
UFC 3-340-02 ‘Structures to resist the effects of accidental explosions’, US Department of
Defense (DoD). Washington, DC, 2008.
Urgessa, G.S., Maji, A.K. ‘Dynamic response of retrofitted masonry walls for blast loading’,
Journal of Engineering Mechanics, 136(7), 858-864, 2010.
Vlassis, A.G., Izzuddin, B.A., Elghazouli, A.Y., Nethercot, D.A. ‘Progressive collapse of
multi-storey buildings due to sudden column loss – Part II: Application. Engineering
Structures, 30(5), 1424-1438, 2008.
Wang, W., Zhang, D., Lu, F. ‘The influence of load pulse shape on pressure-impulse
diagrams of one-way RC slabs’, Structural Engineering and Mechanics, 42(3), 363-381,
2012.
Watwood V.B. ‘Mechanism generation for limit analysis of frames’, Journal of Structural
Division, 109(ST1), 1-13, 1979.
Wisniewski, D., Casas, J. R., Ghosn, M., ‘Load capacity evaluation of existing railway
bridges based on robustness quantification’, Structural Engineering International, 16(2),
161-166, 2006.
Xu, G., Ellingwood, B.R. ‘An energy-based partial pushdown analysis procedure for
assessment of disproportionate collapse potential’, Journal of Constructional Steel
Research, 67, 547–55, 2011.
Xue, B., Le, J.-L. ‘Simplified energy-based analysis of collapse risk of reinforced concrete
buildings’, Structural Safety, 63, 47-58, 2016.
Xue, B., Le, J.-L. ‘Stochastic computational model for progressive collapse of reinforced
concrete buildings’, Journal of Structural Engineering, 142(7), 1-14, 2016.

104
6 Novel proposals for robustness provisions
The aim of this Chapter is to discuss specific practical strategies for the assessment and
realisation of structural robustness, with particular focus on the avoidance of
disproportionate collapse under local damage scenarios in building and bridge structures.
The presented material draws on the main recommendations made by project team
WG6.T2 working under mandate M515 of CEN/TC 250. The Chapter presents a
consideration of both prescriptive and performance based approaches. A novel treatment
of tying forces is presented and discussed in the context of concrete, steel, composite,
timber and aluminium structures. Detailed examples, drawn from the reports of the project
team WG6.T2 are presented to illustrate application of the proposed methodology.

6.1 Building Structures


This section deals with the assessment of multi-storey building structures, with particular
emphasis on a local damage scenario consisting of the sudden loss of a vertical load bearing
member. Besides offering a practical test of structural robustness, enabling the direct
comparison of candidate structural designs, this scenario was previously shown to offer an
upper bound on the response of the building structure in comparison with a scenario in
which the column is damaged by blast loading (Gudmundsson & Izzuddin, 2010).
The two most commonly applied strategies related to robustness design and assessment
are i) tying force methods, and ii) alternative load path methods. An overview of these
strategies, including recent developments, is provided in the following sub-sections.
Focus is given to horizontal tying, which is applicable to the loss of vertical load bearing
members, including columns and walls. While the additional contribution of intact walls and
infill panels is currently considered in the tying force method only simplistically via a
reduction factor, this can be considered effectively using the alternate load path method.

6.1.1 Horizontal Tying Force Strategy


Horizontal tying is recognised in the most recent UFC design code (DoD, 2009) as a means
of bridging over a lost vertical load bearing member. Although this intent is not explicitly
stated in the Eurocodes, EC0 (EN 1990: CEN, 2005) and EC1 (EN 1991-1-7: CEN, 2006),
the consideration of tying under robustness requirements implies an intent to minimize
local damage and to avoid progressive collapse under local damage scenarios. It should
also be noted that the tying rules for frame structures originate from the tying rules for
large panel structures. Within this context, there is a significant discrepancy between the
intensity factors used in the UFC and those used in the current Eurocodes, where a typical
intensity factor for tying via beams is 3.0 and 0.8 for the UFC and Eurocodes, respectively.
While the UFC has taken a fixed level of rotational ductility of 0.2 rad as a realistic value
for typical forms of building construction, the tying force requirements in the Eurocodes
neglect ductility considerations completely, with the intensity factor of 0.8 necessitating
more than 3 times the rotational ductility assumed in the UFC code, which is grossly
unrealistic.
As part of mandate M515 for project team WG6.T2 to enhance the robustness requirements
for the next generation of the Eurocodes, a new rational tying force method has been
proposed by Izzuddin (Izzuddin & Sio, 2022). While this tying force method is based, as a
rational approach, on similar principles as the performance-based alternative load path
method, it is presented within a simplified framework that is much more practical for
robustness assessment and design, making it a suitable replacement for the current
prescriptive tying force requirements in the Eurocodes. Moreover, greater prescription can
be imposed on the new tying force method, such as the prescription of a specific rotational
ductility depending on the construction form and material, thus enabling more
simplification in the application to robustness assessment and design practice.
The new method for horizontal tying (Izzuddin & Sio, 2022) is cast within a simplified
framework, with the explicitly stated aim that horizontal tying is intended to offer a safe

105
bridging catenary/membrane mechanism under a double-span condition in the event of
sudden loss of a column/vertical load bearing member. Importantly, it allows for variable
ductility levels, realistic representation of beams and slabs, and dynamic effects;
moreover, consideration is given to the strength and stiffness requirements from the
surrounding structure to support the horizontal tying forces and the redistributed gravity
loads.
The new approach is formulated in such a way so as to allow the superposition of different
types of load – including floor distributed load, line load and point load – and different
sources of tying within a single floor system – including uniformly distributed reinforcement
in one or two orthogonal directions and tying along a line such as via a beam. The general
formulation is given by (Izzuddin & Sio, 2022):

i  
T    f  P,  ,  in rad  (6-1)
  0.2
where P is the total equivalent load obtained as a superposition from all loads applied to
the double-span beam/floor system, and T is the total equivalent tying force obtained as a
superposition from all active tying forces within the beam/floor system. The remaining
parameters enhance the treatment considered in existing robustness design codes (EN
1991-1-7, 2006; DoD, 2009) as follows:
— if is a tying force intensity factor that depends on the system under consideration. For
reference, this is currently taken as 0.8 in Eurocode 1 (EN 1991-1-7, 2006) and as 3.0
in UFC (DoD, 2009), with the latter widely acknowledged to be more realistic for typical
levels of ductility;
— 𝛼 = 𝛼/0.2 is introduced in the proposed approach to allow for different levels of chord
rotation ductility 𝛼 (rad) in different materials and forms of construction;
— η is a dynamic amplification factor allowing for the influence of ‘sudden’ column/load
bearing member loss; and
— ρ is a reduction factor that allows for such effects as strain-hardening, contributions of
infill walls/infill panels and interaction between tying and flexural actions (taken
conservatively as 1). Further research is needed to establish suitable reduction factors
for such effects, with progress already made on representing the interaction between
tying and flexural actions (Izzuddin, 2022).
The above Equation (6-1) provides the basis for checking the adequacy of the provided
equivalent tying force T (as assembled from different active components within the affected
beam/floor system) to resist the equivalent load P (as assembled from different loads
applied to the system) for the maximum normalised rotational ductility 𝛼 of the system.
This clearly requires knowledge of the rotational ductility α, which typically depends on the
construction material and structural form. Table 1 presents the parameters if, T and P, as
required in Equation (6-1), for selected systems subject to uniformly distributed loading,
specifically i) tying via beam (symmetric mode), ii) two-way tying via floor slab (with
distributed slab reinforcement), and iii) one-way tying via floor slab (with distributed slab
reinforcement, assuming torsional edge restrains). Full details, including other types of
loading and sources of tying, can be found in (Izzuddin & Sio, 2022).
For consistency with alternative load path methods, dynamic amplification is introduced in
the proposed tying force method through factor η, which establishes an equivalent
amplification of the loading under sudden loss of a column or vertical load bearing member.
The development of load resistance with tying is typically linear as a function of the chord
rotation after the attainment of the minimum chord rotation 𝛼𝑚𝑖𝑛 , as illustrated in Figure
6-1. In the absence of information on other resistance mechanisms before the development
of tensile catenary/membrane action, the most realistic value for the dynamic amplification
factor is therefore η = 2. Notwithstanding, a refinement of η can be obtained if the nature
of the response preceding the attainment of full tensile catenary/membrane action is

106
known. The most typical case relates to flexural action at lower levels of deflection/chord
rotation, as illustrated in Figure 6-1, where it may be assumed that the maximum flexural
resistance Pf is achieved at relatively small deflections. Using energy balance principles
(Izzuddin et al., 2008; Izzuddin, 2010), a refined dynamic amplification factor can be
readily obtained according to an explicit expression (Izzuddin & Sio, 2022).

Figure 6-1. Development of tying and flexural resistance

Source: Izzuddin & Sio, 2022

An important factor affecting the development of horizontal tying forces following the loss
of a vertical load bearing member is the planar restraint offered by the structure
surrounding the affected floor system, both in terms of stiffness and strength. While the
planar strength has to be checked against the ultimate tying force capacity, the planar
stiffness depends on the rotational capacity of the floor system, with a lower planar
stiffness demanding a greater rotational capacity before the development of
catenary/membrane action. For a given planar stiffness from the surrounding structure, a
minimum chord rotation 𝛼𝑚𝑖𝑛 is established in the new method (Izzuddin & Sio, 2022) for
different structural tying systems. Provided that the rotational capacity 𝛼 exceeds 𝛼𝑚𝑖𝑛 , full
catenary/membrane action can be developed; otherwise, other resistance mechanisms
(e.g. flexural/compressive arching action) should be considered.
Another important factor relating to the interaction with the surrounding structure is the
dynamic amplification of the redistributed gravity loading. In this respect, a static
redistribution is unsafe for a sudden vertical member loss, while a redistribution based on
the dynamic load amplification factor η used for the deformation demands can be grossly
conservative, due to difference between the distributions of load and inertia forces. A more
realistic amplification of the redistributed gravity loading to the surrounding structure can
be obtained accounting for the difference, as given in Table 6.1 for the selected cases, with
full details covering other cases presented in (Izzuddin & Sio, 2022).

107
Table 6-1. Tying parameters and redistributed load amplification for selected 1D/2D systems

Tying via beam Two-way tying via floor One-way tying via floor

fy
F F f f
fx fx
   H
2H 2L
2L
fy 2L

Intensity
factor: 2.5 3.125 3.125
if

Equivalent
 L fH
tying force: F fx H  f y L  
T H 2

w w
q
Loading
2H H
2L 2L

Equivalent wLH
load: qL wLH
P 2

Redistribut
ed load
0.25  0.75 0.3056  0.6944 0.3056  0.6944
amplificati
on
Source: Izzuddin & Sio, 2022

6.1.2 Application of different structural Systems/ Materials

6.1.2.1 Concrete structures


Several applications of tying systems are available in the scientific literature for reinforced
concrete frames. Distributed or concentrated tying systems can be adopted to avoid
disproportionate collapse under local damage scenarios. In Belletti et al. (2019) a database
of experimental tests on peripheral ties in reinforced concrete frames sub-assemblies is
illustrated, while in the report by CEN/TC250/WG6 Project Team WG6.T2 “Experimental
Database and Validation of Analytical Models for Tie Force and Alternative Load Path
Methods for RC Structures” (Belletti et al. 2021) a more complete database, which includes
also experimental tests on internal two-way tying systems concentrated in beam and
peripheral and internal tying systems distributed in slabs, is reported. The results collected
in the database allow to evaluate the rotational ductility of the investigated members in
order to provide useful information for the application of the tie force method in the
engineering practice.
The capacity of tying systems is experimentally observed by simulating the Column Loss
Scenario with static or dynamic tests. The column loss scenario can be applied in different
positions with respect to the building plan – such as interior, peripheral and corner

108
locations. Most of the experimental tests have been carried out on structural sub-
assemblies of RC frames, by analysing concentrated tying systems; few experimental tests
on full‐ scale RC buildings have been conducted. Few experimental tests on the
dependency of robustness performance on the position in-elevation of the column loss
scenario in multi-story buildings are available. Experimental tests on RC sub-assemblies
deal with the tie’s capacity dependency on brittle and ductile failure modes - in beams,
columns, beam-to-columns joints and slabs -, compressive membrane actions - which
enhances the flexural and shear resistance -, tensile membrane action - which activate the
catenary stage -, tie anchorage and tie continuity, ductility and dynamic effects.
Experimental tests on full‐ scale structures allows to appreciate the tie’s capacity
dependency on boundary conditions, activation of Vierendeel actions and internal forces
redistribution. No data are available on experimental test on precast structures with tying
systems and pinned connections. Few experimental investigations on tying systems in
precast moment resisting frames are available. Therefore, extension of the concepts
applied to cast-in-situ concrete structures may be carefully applied to precast structures.
For the robustness verification of precast structures, experimental tests are highly
recommended to evaluate the effective tying resistance and the equilibrium and
compatibility conditions to be respected. The compatibility requirements are particularly
important in the case of precast structures realised with dried connections.
The tie method proposed by Izzuddin & Sio (2022) has been validated in Belletti et al.
(2021) and Ravasini at al. (2021) by comparing the analytical results with the experimental
results obtained from the tests reported in the database. The rotational ductility has been
obtained from the average value of the rotational ductility observed from experimental
test. It must be noted that for distributed tying systems or for coupled tying systems -
with concentrated tying reinforcement in the beam and distributed tying reinforcement in
the slab, the population of the tested specimens is not representative for a reliable
estimation of the ductility level and that further experimental results are required. The
ratio between the experimental resistance and the resistance calculated according with the
Tying force method has been evaluated for each analysed specimen, Figure 6-2. These
latter resistances are evaluated in correspondence of the assumed rotational ductility. The
total equivalent tying force is calculated as a function of the tying reinforcement area and
the yield strength of steel. The average value, the standard deviation and the Coefficient
of Variation (CoV) of this ratio have been evaluated for each group of sub-assemblies (1D-
beam, 2D-beam, beam-slab and flat-slab assemblies).

Figure 6-2. Experimental and analytical tie resistance at the catenary stage evaluated at the
assumed ductility level
Vertical load

Experimental resistance
Analytical resistance Vertical load

Displacement
corresponding to the
Vertical displacement
assumed chord rotation

Vertical displacement

The main hypotheses assumed for the validation and the main results are reported in the
following:
— The tie resistance is calculated by adopting the yield strength of steel;
— 1D-beam sub-assembly: a chord rotation equal to equal to 0.2 rad is assumed for
beams characterised by a ductile failure and plastic hinges formations at beam ends.
The average value of the ratio between the experimental and the analytical resistance

109
is equal to 1.02; the standard deviation is equal to 0.3 and the coefficient of variation
is equal to 29.4%;
— For beams characterised by values of span‐to‐depth ratio lower than 9, the failure mode
is usually governed by brittle mechanisms and lower ductility levels must be adopted.
Indeed, the flexural resistance enhanced by compressive arch action results higher
than the resistance at the catenary stage;
— 2D-beam sub-assembly: the capacity of a tying system composed by two intersecting
beams is obtained by adopting the principle of superposition by adding the contribution
of each individual beam. The maximum displacement at failure is achieved in
correspondence of the minimum ductility of the two intersecting beams;
— 2D-beam-slab sub-assembly: the capacity of the tying system composed by 2D-beams
and the slab is obtained by adopting the principle of superposition by adding the
contribution of each individual beam and the contribution of the slab. In the case of
two intersecting beams and slab sub‐assembly of frame systems, if the non‐linear
response is governed by ductile flexural mechanisms a chord rotation value equal to
0.2 may be assumed. If the non-linear response is governed by shear failure or
punching failure a lower ductility level must be assumed. It must be noted that the
population of the tested specimens available in the database is not representative for
a reliable estimation of the ductility level and that further experimental results are
required. The average value of the ratio between the experimental and the analytical
resistance is equal to 1.37; the standard deviation is equal to 0.14 and the coefficient
of variation is equal to 10%;
— In continuous flat slabs, supported by columns, the non‐linear response is usually
governed by the punching shear. The maximum ductility must be carefully evaluated.
Integrity reinforcement can enhance the post‐punching capacity both in term of
resistance and ductility; in general, the maximum chord rotation should be lower than
or equal to 0.1 rad. It must be noted that the population of the tested specimens
available in the database is not representative for a reliable estimation of the ductility
level and that further experimental results are required.
— Tie method in precast frame systems has been rarely applied. Few numerical data on
the coupled effect of pinned beam-to-columns connections and concentrated ties are
reported in literature (Ravasini et al. 2021a). Few experimental investigations on tying
systems in precast moment resisting frames are available.
In the Report of the Project Team WG6.T2 some examples of application of the tie force
method to RC Systems Under Interior Column Loss are provided. In particular, the cases
of concentrated ties and distributed ties have been analysed for frames having an in-plane
distance between columns equal to 8.1m. The tying reinforcement adopted in the example
is assumed continuous and well anchorated.
In the first example, a precast moment resisting frame is analysed. The tying reinforcement
- concentrated along the beam supporting the floor slab and constituted of 3 Ø24 + 4 Ø26
at the top and 5 Ø24 at the bottom longitudinal continuous bars – allows to verify the
capacity of the structure for a distributed load, applied to the beam and evaluated in the
accidental load combination, equal to 63.34 kN/m. Of course, additional verifications, not
reported in the example are required to avoid brittle failures occurring before the formation
of the catenary mechanism. Therefore, appropriate resistance and compatibility controls
of the connections between the topping slab and the hollow‐core slabs and of the
connections between the transversal beam and the hollow‐core slab have to be carried out.
Furthermore, appropriate check of the hollow‐core slabs resistance and of shear resistance
of the beams are required.
In the second example, the RC frame is realised by cast in situ members and the tying
system is distributed in the slab. The total equivalent tying force is calculated as a function
of the continuous bottom (sagging) reinforcement (Ø14/250) and the continuous top
(hogging) reinforcement (Ø16/300). Since the contribution of the slab has been considered

110
under applied surface load, a chord rotation ductility of α = 0.1 rad is assumed. It results
that the slab offers a safe bridging membrane mechanism under the event of sudden loss
of the interior column when a distributed load, corresponding to the accidental load
calculation, equal to 10.2 kN/m2 is applied. Of course, additional verifications, not reported
in the example are required to avoid brittle failures occurring before the formation of the
catenary mechanism. Therefore, an appropriate check of the punching and post‐punching
resistances is required.

6.1.2.2 Steel and Composite structures


The properties and behaviour of rolled steel sections are well known and thoroughly
investigated by various extensive (European) studies (EUR 28906, 2017). Consequently,
the results of this type of studies are only valid for steel grades and production methods
that comply with the various European product standards. Steel members can be
recognised as ductile under the various load types if instability effects can be avoided by
proper means. It is also known that mostly a lack of robustness can be mainly attributed
to the joints which turn out to be critical (Knoll & Vogel, 2009) or the zones immediately
adjacent. In that way, the proposed Tying Force Strategy (TFS) out of section 6.6.1
perfectly match the behaviour of steel elements as in the underlying model; all plasticity
is concentrated at the locations of the nodes, which in general coincides with the location
of the assemblies. This system property can be even more explored so that with relatively
simple numerical aids, a complete robustness assessment can be obtained (Molkens,
2021).
So-called mild steels with low to modest yield strength exhibit an increase of the strength
(stress) with increasing deformation (strain), Figure 6-3. This property is used in regular
steel design where the following criteria should be met: A × f y < Anet × fu with A = the
surface area of a member without holes, fy (or fy,nom) = the yield strength, Anet = reduced
surface area with bolt holes and ft (or fu,nom) the tensile ultimate strength. Similar criteria
can be stated for connections designed for higher strength than one of the beam elements.

Figure 6-3. Engineering stress-strain relationship for steel (Yun & Gardner, 2017)

Source: Yun & Gardner, 2017

Out of analysis of the data from some experiments (Dinu, Marginean, & Dubina, 2017) and
(Demonceau et al., 2010)), even in combination with large deformations, the tested
connection can still transfer some moments. Favourable effects of strain hardening and
possible combinations of moment and arch- or catenary effects can be implemented by the
 factor out of the proposed Eq. (6-1). A European valorisation project involving 14
European partners coordinated by the University of Liège is ongoing and will be released
in 2022. Outcomes of this project will be of high importance to address the behaviour of
steel and composite structures and, in particular, to ensure a ductile behaviour and the
possibility of activating alternative load paths (Demonceau et al, 2022). For the time being,
designers can use data out of (UFC 4-023-03, 2016) with the proposed chord rotation α =

111
0,20 rad, besides concrete also valid for steel and composite. For cold-formed steel
components, it should be proven that the system can carry the required longitudinal,
transverse, and peripheral tie strength before applying this chord rotation.
In some instances, partial buckling of the member zones adjacent to connections may be
permissible provided that the changed load arrangement (pure tension in case of full
catenary action instead of shear, normal force, and moments) can be safely transmitted to
the rest of the structure. Ductile failure modes of the joints are, under any circumstance,
obligatory when designing for robustness except in case of full-strength joints. In practice,
this means that the failure of bolts, welds or other fasteners should be avoided (Demonceau
et al, 2022).
The design of full-strength joints should account for overstrength, which is also typical for
mild steels, the data out of prEN 1998 can be used, see Table 6-2. One way to achieve it,
based on the same philosophy as for strain-hardening, is the following: the nominal or
code specified values should be used to determine the resistance of the joint. An
assessment should be carried out to check if the beam capacity based on the mean values
including an overstrength or randomness coefficient γrm is lower than the resistance of the
joint. The designer should be aware that this will lead to rather massive and mostly non-
economical solutions.

Table 6-2. Recommended value of γrm

Steel Grade 𝜸𝒓𝒎

S235 1,45

S275 1,35

S355 1,25

S460 1,20
Source: prEN1998, not yet published

Significant research activities have been conducted in the past decades to investigate the
behaviour of steel and composite framed structures subjected to sudden column removal
in the framework of RFCS (Research Fund for Coal and Steel) projects involving different
European research institutions. The main outcome coming from these projects are reflected
in (Demonceau et al, 2022).
Information about robustness related items of structures made of stainless steels is even
more rare to find. However, high ductility of these steels (strains of 20 up to more than
50 % are usual) associated to substantial strain hardening leads to the conclusion that a
design following the rules for carbon steel will result in a (over) conservative design. Proper
methods considering a Ramberg-Osgood constitutive behaviour law like the continuous
strength method out of (SCI, 2017) can enhance material use and reduce related costs.
As stated before, it is also possible to adapt the design from a conceptual point of view to
avoid possible problems; some sound engineering principles can be followed:
— To promote the activation of ductile structural components in the structure and,
accordingly, to avoid brittle failure modes, with a specific attention to be paid to the
welds (the use of full strength welds is recommended);
— To ensure links between the elements using the tying approach according to the
proposed TFS;
— Optimise the design making use of strain-hardening and combined actions if this can
be supported by tests or reliable data.

112
Some application examples can be found in the Report of the Project Team WG6.T2,
starting with a continuous steel beam of a 2D frame. In a second example, the spacers
typically used to limit buckling length and avoid lateral-torsional buckling are also
activated. In a third simulation, the reinforcement mesh of the compression layer is used
to provide additional contribution for robustness. Under normal conditions, there is no
contribution of the compression layer (except superimposed dead load), but in the
accidental situation, the topping behaves as a membrane and no requirements are set to
the steel structure. The fourth example makes use of all components, topping, beams and
spacers. For that reason, all these components should satisfy robustness requirements. In
the case of existing structures, an approach whereby new additional elements serves as
robustness provisions can be the most straightforward approach to obtain or enhance
robustness. In some extend this is reflected in example three.

6.1.2.3 Timber structures


The tie method proposed by Izzuddin & Sio (2022) has been validated in Martinelli and &
Izzuddin (2022) as regards the timber post-and-beam structural typology, the only one at
the moment for which some experimental data on rotational ductility under column loss
scenario is available.
The newly proposed tying force formulation given in Eq. (6-1), depends on the parameter
𝛼̅ = 𝛼/0.2 that accounts different levels of chord rotation ductility 𝛼 (rad) in different
materials and forms of construction.
With reference to post-and-beam timber structures, the rotational ductility is strongly
influenced by the type of connection between beams and columns. Although the rotational
ductility data available in the literature are rather limited and refer to a single structural
typology of timber constructions (i.e. post-and-beam structures), it is still considered
important to analyse here the data presented in the literature. Based on the experimental
data reported by Masaeli et al., (2020) and by Lyu et al., (2020), the ultimate rotations
for the different types of beam-column connectors are reviewed in this section.
Three types of commercially available beam-to-column connectors, currently used in mass
timber buildings, were investigated by Lyu et al., (2020). A description of the four
connectors is also given in the report by CEN/TC250/WG6 Project Team WG6.T2
“Considerations on the rotational ductility in timber post-and-beam structures” (Martinelli
2021). These connectors were designed as shear connectors, not moment resisting ones,
assuming that the horizontal stability of the structure is ensured by shear walls and cross-
bracing elements instead of frame actions. The three types of beam-to-column connectors
were not specifically designed for robustness. The scaling effects of these three types of
connectors on the moment and shear responses were investigated in Masaeli et al., (2020)
where full-scale and ¼ scale connectors were compared. The collapse response of the three
types of connectors under a quasi-static column removal scenario (i.e. push down test)
was investigated in Lyu et al., (2020) adopting a ¼ scale prototype.
A fourth connector was proposed by the same research group (Lyu et al., 2020) and
designed to resist the loss of a column through catenary action. The performance of all
four types of connectors was studied by Lyu et al., (2020). All connectors were designed
based on the specification given in EN 1995-1-1 (2004) to sustain the same factored design
shear force of a representative building. The factored design shear force of the beam-to-
column connectors under medium-term actions was equal to 183 kN for the full-scale
connectors.
Table 6.3 reports the ultimate rotations for connectors Types 1-3 obtained from bending
tests at full-scale and at ¼ scale. Data reported in Table 6-3 are in good agreement with
those obtained from push-down tests at ¼ scale (Table 6-4). All commercial connectors
(Types 1-3) provided enough rotation for the catenary action to either develop or start
developing. Only connector Type 1 shows a limited rotation ductility that prevents a full
development of catenary action during the push-down test. The other types of connectors
(Types 2-4) show a good rotational ductility that allow a full development of catenary action

113
during the push-down tests. The connector Type 4 (double plate connector) proposed by
Lyu et al., (2020) show a higher capacities and higher rotational ductility compared to the
other connectors. Connector Type 4 represents a potential solution to improve the
robustness of post-and-beam timber buildings.

Table 6-3. Ultimate rotations for connectors Types 1-3 from bending tests at full-scale and at ¼
scale

Type 1 Type 2 Type 3

 (rad) Full-scale test 0.12 0.087 0.16

 (rad) ¼ scale test 0.070 0.23 0.16


Source: Masaeli et al., 2020

Table 6-4. Ultimate rotations for connectors Types 1-4 from push-down tests at ¼ scale

Type 1 Type 2 Type 3 Type 4-1 Type 4-2 Type 4-3

 (rad) ¼
0.07 0.17 0.15 0.21 0.21 0.17
scale test
Source: Lyu et al., 2020

In the background report of the Project Team WG6.T2 (Martinelli 2021) a preliminary
validation of the tie force method to post-and-beam timber systems under interior column
loss is provided. In the Report of the Project Team WG6.T2 one example of application of
the tie force method to post-and-beam timber structure is given.
In the example, a 4 storeys 5 × 5-bays post-and-beam timber structure is considered
(§6.3.3). The structure is used as offices (building category B according to BS EN 1991-1-
1, 2002) and falls in consequence class 2a according to BS EN 1991-1-7 (2006). Span
lengths along x- and y-direction are equals to 8 and 6 m, respectively, while the inter-
story height is equal to 3.6 m. Continuous multi-storey glued laminated columns are
adopted. Glulam is used for both primary and periphery beams, labelled as B1 and B2, and
for continuous multi-story column labelled as C1. CLT slabs 295 mm thick are adopted.

6.1.2.4 Aluminium Structures


The design of aluminium structures for robustness can be executed in a very similar way
as for steel. The designer should be, however, aware about:
— Only a few grades are suitable for structural works. The in Europe, usable wrought and
cast alloys can be found in Tables 3.1a and 3.1b of CEN (2007).
— The material is strain hardening with a Ramberg-Osgood constitutive relationship, see
Figure 6-4.

114
Figure 6-4. Engineering Typical engineering stress-strain curve for aluminium alloys

Source: Yun et al., 2021

6.1.3 Alternative load path strategy


The alternative load path methods offer a more performance-based approach to the
robustness assessment of structures compared to tying methods, allowing for other
resistance mechanisms such as flexural and compressive actions as well as for a more
accurate representation of the interaction between the affected part of the structure and
the surrounding parts. Focus is given here to the application of alternative load path
methods to multi-storey building structures subject to sudden column loss as a local
damage scenario.
A simplified multi-level robustness assessment framework was proposed by Izzuddin et al.
(2008) for multi-storey buildings subject to sudden column loss, which can be applied at
various levels of structural idealisation, and which combines the influence of redundancy,
ductility, energy absorption and dynamic effects into a single measure of robustness,
namely the pseudo-static capacity. To illustrate the multi-level aspect of this method, with
reference to the building structure subject to sudden column loss as shown in Figure 6-5,
it is the ability of the upper floors to sustain the ensuing dynamic deformations that
determines whether disproportionate (progressive) collapse can be avoided. Accordingly,
alternative reduced models may be considered at various levels of structural idealisation,
as illustrated in Figure 6-6, depending on the regularity of the building in terms structure
and loading (Izzuddin et al., 2008). At the first level of model reduction, consideration may
be given to the affected bay of the multi-storey building (Figure 6-6a), with appropriate
boundary conditions to represent the interaction with the surrounding structure. If it is
established that the surrounding columns can resist the redistributed load, further model
reduction may consider only the floors above the lost column where deformation is
concentrated (Figure 6-6). If additionally the affected floors are identical in terms of
structure and loading, the axial force in the columns immediately above the lost column
becomes negligible, and a reduced model consisting of a single floor system may be
considered (Figure 6-6c). Finally, ignoring planar effects within the floor slab, individual
beams may be considered at the lowest level of model reduction (Figure 6-6d), subject to
appropriate proportions of the gravity load.
Once the extent of the structural model is defined, it is used to establish the ability of the
locally damaged structure to resist the applied gravity loading, allowing for dynamic
effects, without further failure. Consideration should be given to the factored gravity
loading and to the ensuing dynamic effects due to sudden local damage which can be
assessed by means of static or dynamic analysis methods. In all cases, the robustness limit
state should be concerned with the ability of the locally damaged structure to maintain
integrity through the explicit consideration of redundancy, energy absorption capacity and
ductility limits under dynamic conditions (Izzuddin et al., 2008). This can be based on
ensuring that no component in the structure outside the locally damaged region exceeds
its deformation or strength limit, as appropriate. Allowance can also be made for successive
component failures (Izzuddin, 2010), provided these do not lead to collapse outside the
locally damaged region due to sufficient residual dynamic strength in the surrounding

115
structure. While some types of structure may be assessed for robustness using linear
analysis methods, large inelastic deformations would be allowed and expected in typical
forms of construction, such as steel-framed and reinforced concrete buildings, which
necessitate the use of nonlinear analysis. Moreover, while nonlinear dynamic analysis offers
the most accurate representation of the structural response under sudden column loss,
focus is given here to nonlinear static methods combined with simplified dynamic
assessment as a more practical approach for application in robustness design and
assessment practice.
Nonlinear structural analysis offers major advantages compared to linear analysis,
particularly for structures in which components fail by fracture well after the development
of significant plastic deformation. In this type of analysis, both geometric and material
nonlinearity should typically be considered, allowing for the modelling of large
displacements, structural instability and inelastic material response (e.g. yielding of steel,
cracking of concrete). Depending on the sophistication of the adopted models, nonlinear
phenomena preceding component failure/structural collapse can be represented to various
degrees of accuracy, including the flexural strength, compressive arching action, tensile
catenary/membrane action, diaphragm action of slabs/walls, etc. In this respect, the use
of nonlinear analysis in robustness assessment allows the consideration of the structure at
relatively large deformations thus exceeding conventional strength-based norms, provided
the deformation capacities at which components fail by fracture are not exceeded.
Accordingly, the robustness limit state requires the structure subject to a sudden local
damage scenario to exhibit component deformations that are within their respective
deformation capacity at fracture, collectively defining a so-called ductility limit on the
maximum dynamic deformations of the locally damaged structure (Izzuddin et al., 2008).

Figure 6-5. Multi-storey building subject to sudden column loss

Source: Izzuddin et al., 2008

116
Figure 6-6. Sub-structural levels for robustness assessment

Source: Izzuddin et al., 2008

There are numerous types of elements that can be used for nonlinear structural analysis,
the review of which is outside the current scope. Some indicative references on nonlinear
analysis of structural components are as follows:
— 1D beam-column elements of the fibre-type can be used to model steel (Izzuddin &
Elnashai, 1993) and reinforced concrete (RC) (Izzuddin & Lloyd Smith, 2000) frames
with reasonable accuracy.
— 2D shell elements can be used to model RC and composite floor slabs (Izzuddin et al.,
2004), allowing for the development of membrane action, while an approximate grillage
representation of the floor slabs can also be used for a conservative assessment of
robustness (Zolghadr Jahromi et al., 2013).
— Masonry infill panels can be modelled at different levels of sophistication, including as
equivalent 1D struts (Farazman et al., 2013), simplified macro-elements (Minga et al.,
2020), and sophisticated 3D meso-scale models (Xavier et al., 2015). Previous work
has shown that the incorporation of infill panels in the structural model can lead through
vertical diaphragm action to a significant enhancement of robustness, with reduced
requirements on the deformation limit of the structure.
— Discrete elements may be used to model RC concrete joints (Favvata et al., 2008) and
steel connections (Fang et al., 2013), including in the latter case component-based
methods (Steenhuis et al., 1998) that are available in EC3. It is important to note that,
in application to robustness scenarios, the joint models must capture the interaction
between bending and axial actions not only in terms of the force-deformation
relationships but also in terms of the assessment of the deformation capacity; this goes
beyond the specific requirements of joint ductility under seismic action, where the axial
action in the joints is typically neglected.
For nonlinear analysis models which do not account for the component response following
the initiation of fracture failure, the robustness assessment can be based conservatively
on the initiation of first-component failure. It is also worth highlighting that nonlinear

117
analysis models of different sophistication may be applied at the various levels of structural
idealisation highlighted before, where lower-level models are often sufficient for a
preliminary conservative assessment of structural robustness (Izzuddin et al., 2008;
Zolghadr Jahromi, 2013).
Focusing on nonlinear static analysis methods for robustness assessment of building
structures subject to sudden column loss, this involves a pushdown scenario utilising
gravity loading, leading to a static load-deflection response of the locally damaged
structure, with a typical example shown in Figure 6-7. Such a pushdown nonlinear static
response can be generated for the full structure or substructures, as discussed previously,
making use of detailed nonlinear finite element models or simplified mechanics-based
models (Izzuddin et al., 2008), though it is important to highlight the potential inaccuracy
of the latter models and their typical inability to represent well the interaction with the
surrounding structure.
The nonlinear static pushdown response can be used to recover the maximum dynamic
deformation, resulting from the sudden nature of column loss, as that corresponding to an
amplified static load with a factor η, as illustrated in Figure 6-5. This dynamic amplification
factor does not need to be assumed, but it can be easily recovered using a simplified
dynamic assessment approach based on an energy balance concept (Izzuddin et al., 2008),
which assumes that the pushdown response is governed by a dominant deformation mode,
consequently leading to a close similarity between the effects of sudden local damage and
loading that is applied suddenly to the locally damaged structure (Izzuddin, 2010). The
simplified dynamic assessment approach is illustrated in Figure 6-6, where for any level of
applied gravity loading (λiP0), the corresponding dynamic displacement (ud,i) can be
obtained from the nonlinear static pushdown response by equating the work done by the
suddenly applied load (rectangular hatched area) to the internal energy absorbed by the
structure (hatched area under nonlinear static curve). The check for the robustness limit
state at a specific level of gravity loading (λiP0) would then consider whether the dynamic
displacement (ud,i) is within the ductility limit of the locally damaged structure, or
alternatively whether all the components of this structure are within their respective
deformation capacity.

Figure 6-7. Nonlinear static pushdown response

Source: Izzuddin, 2010

118
Figure 6-8. Simplified dynamic assessment and pseudo-static response (Izzuddin et al., 2008)

Source: Izzuddin et al., 2008

The simplified dynamic assessment approach provides a simple transformation of the


nonlinear static pushdown response to a maximum dynamic response, so-called pseudo-
static response (Figure 6-8c), as it inherits similar characteristics to the static response,
including the assembly of resistance contributions from different parts of the locally
damaged structure (Izzuddin et al., 2008). This transformation implies that the pseudo-
static resistance at a specific level of dynamic deformation is the average of the static
resistance up to this level of displacement. This approach does not require the explicit
determination of the dynamic amplification factor, though this can be easily recovered as
the ratio of the static resistance to the pseudo-static resistance at a specific level of
displacement.
Unlike nonlinear dynamic analysis, the use of nonlinear static analysis with the simplified
dynamic (pseudo-static) assessment enables the consideration of robustness at different
levels of gravity loading without the need for re-analysis, as the relevant margin of the
applied loading (λiP0) to the resistance (Pf) at the robustness limit state is already available
from the pseudo-static response (Figure 6-8c). However, besides the assumption of a
dominant deformation mode, the applicability of the standard pseudo-static assessment to
successive component failures requires the same mode to persist after component
failure(s). Further information on the treatment of successive component failures in
robustness assessment using the energy balance concept can be found in the work of
Izzuddin (2012).

6.1.4 Examples of the application of the novel proposals


The following examples are reproduced from the final report of PT WG6.T2 on Robustness
Rules in the Material Related Eurocode Parts. The calculation scheme employed for the newly
developed tying force method, which represents an assessment procedure, is as illustrated in
Figure 6-9.

119
Figure 6-9. Performance indicators

6.1.4.1 Reinforced Concrete Structure Example


A frame is realised by precast members. The plan view is depicted in Figure 6-10. The
dimensions of beams and columns are:
— Beams: height equal to 60 cm and width equal to 60 cm;
— Columns: height equal to 80 cm and width equal to 80 cm.
The diaphragm is realised by hollow-core slab and by a topping slab 5 cm thick. Continuous
beams support the diaphragm.
In this example it is assumed that the contribution of the topping slab, to the tie force
assessment, is neglected. Therefore, the tie reinforcement is concentrated in the beam
supporting the hollow-core slabs, which behave as double-span beams. The continuous
longitudinal reinforcements, which is used as tie reinforcement in transversal beams are:
— Top reinforcement: 3 Ø24 + 4 Ø 26
— Bottom reinforcement: 5 Ø24
— As = 5739.92 mm2

120
Figure 6-10. Plan view of the emulative precast RC frame under interior column loss

8100 8100 8100 8100 8100

1 2 3 4 5 6

8100

D
Floor type
8100

C
8100

Beam
B
Column
8100

Y
A
X Plan view (5 x 4 bays)

For both dead and live loads, the values are reported in Table 6-5.

Table 6-5. Applied loads

Load types Magnitude Units

DL: beams 8.83 kN/m

DL: walls 0.80 kN/m2

DL: false floor 0.50 kN/m2

DL: equipment 0.50 kN/m2

DL: hollow-core slab 2.78 kN/m2

DL: topping 1.25 kN/m2

LL: Live Load 3.00 kN/m2

The loads, reported in Table 6-5, are used for the calculation of accidental load
combination. The uniformly distributed load q, acting on the transversal beam, is given as:

q = 8.83 + 8.1 · (0.8 + 0.5 + 0.5 + 2.78 + 1.25) + 0.3 · 8.1 · 3 = 63.34 kN/m (6-2)

The calculation of the equivalent tying force T and load P are reported in Table 6-6.

Table 6-6. Tying parameters for double-span beams

Equal spans q

121
F F

2L

Intensity factor: if 2.5

𝑇 = 𝐴𝑠 ∙ 𝑓𝑦𝑘 = 5739.92 ∙ 450


Equivalent tying force: T F
= 2582.96 𝑘𝑁

𝑃 = 𝑞 ∙ 𝐿 = 63.34 ∙ 8.1
Equivalent load: P qL
= 513.07 𝑘𝑁

Tie capacity verification


As a first demonstration of the tie force verification, the dynamic amplification factor and
the reduction factor are taken as  = 2 and ρ = 1, respectively.
The normalized chord rotation α
̅ is evaluated by assuming a chord rotation ductility equal
to α = 0.2 rad:

𝛼 0.2 (6-3)
𝛼̅ = = =1
0.2 0.2

The adequacy of tying force is evaluated in the following:

2.5 (6-4)
𝑇 = 2582.96 𝑘𝑁 ≥ 2 ∙ 1 ∙ ( ) 513.05 = 2565.25 𝑘𝑁
1

Therefore, for the assumed level of chord rotation ductility and applied vertical loads, the
transversal beam offers a safe bridging catenary mechanism under the event of sudden
loss of the interior column.
Appropriate checks of the connections between the topping slab and the hollow-core slabs
and of the connections between the transversal beam and the hollow-core slab are
required. An appropriate check of the hollow-core slabs capacity is required. An
appropriate check of the shear capacity of the beam is required.

Strength and stiffness requirements of surrounding structure


The surrounding structure must provide the necessary strength to resist the tying forces
induced in any double-span beams and the horizontal floor system. The total pull-in
displacement u must be limited to the following value:

𝐿1 𝑑𝑒𝑓𝑓
2
𝐿1 + 𝐿2 (6-5)
𝑢≤ (𝛼 − ) ( )−𝛿
2 𝐿1 𝐿2

For the evaluation of the displacement u the horizontal force provided by the tie, equal to
F=2582.96 kN, is applied to the topping slab, 5 cm thick, by neglecting the stiffness
contribution provided by the frame and by assuming a reduced value of the modulus of
elasticity for concrete in order to take into account for cracking effects, see Figure 6-11.
Therefore, the lateral displacement has been calculated by accounting for the in-plane
shear and flexural deformation of the topping RC slab.
The lateral displacement to the left uL and to the right uR of the double-span beam results
respectively equal to uL =5.88 mm and uR =2.35 mm.
The total displacement results equal to:

122
𝑢 = 𝑢𝑅 + 𝑢𝐿 = 8.23 𝑚𝑚 (6-6)

The elastic extension of the double-span beam under F, here calculated by neglecting
tension stiffening effects, is given in the following:

450 (6-7)
𝛿 = 2 ∙ 𝐿 ∙ 𝜀𝑦 = 2 ∙ 8100 ∙ = 36.45 𝑚𝑚
200000

The verification of the stiffness of the surrounding structures is provided in the following,
being deff equal to 495 mm:

8100 495 2 8100 + 8100 (6-8)


𝑢 = 8.23 𝑚𝑚 ≤ (0.2 − ) ( ) − 36.45 = 119.8 𝑚𝑚
2 8100 8100

Figure 6-11. Lateral stiffness provided by the topping slab of 5 cm

8100 8100 8100 8100 8100

1 2 3 4 5 6

E
8100

D
Floor type
8100

F F
C
8100

Beam
B
Column
8100

Y
A
X Plan view (5 x 4 bays)

Resistance to redistributed vertical gravity loads


The use of the previous full amplification, η, would be conservative, as this assumes inertia
forces to have the same spatial distribution as the actual gravity loading. Instead, the
amplification factors can be established for the dominant load according to its type and the
system under consideration.

123
Table 6-7. Amplification of redistributed vertical gravity loading for double-span beams

Equal spans
F F

2L

Redistributed gravity
load amplification
q
0.25 + 0.75 ∙ 𝜂 = 0.25 + 0.75 ∙ 2 = 1.75

Minimum rotational ductility for activation of tying


The minimum level of ductility αmin required for the activation of tensile catenary action is
given in the following:

𝑑𝑒𝑓𝑓 2(𝑢 + 𝛿) 𝐿2
𝛼𝑚𝑖𝑛 = +√ ( ) (6-9)
𝐿1 𝐿1 + 𝐿2 𝐿1

The calculation yields:

495 2(8.24 + 36.45)


𝛼𝑚𝑖𝑛 = +√ ∙ 1 = 0.135 𝑟𝑎𝑑 (6-10)
8100 2 ∙ 8100

with u and δ as previously defined.


Since α > αmin, the use of the tie method is justified.

124
6.1.4.2 Steel Structure Example
A frame is realised by steel members. The static system is as shown in Figure 6-12.
Load assumptions
Self-weight: gk = 0.56 kN/m (assume an IPE360 section 57.1 kg/m A-M catalogue,
width 170 mm and height 360 mm)
Dead load: pk = 1.5+8.1/2∙(0.8+0.5+1.25+2.78+0.5) = 25.11 kN/m (façade +
partition walls, false floor, topping, hollow cores and false ceiling with
equipment)
Live load B: qk = 8.1/2∙3 = 12.15 kN/m (office building EN 1991-1-1; Ψ0 = 0.70, Ψ1
= 0.50 and Ψ2 = 0.30)
Service load: q = 0.56+25.11+12.15 = 37.82 kN/m (Eq. 6.14 EN 1990)
Design load: qd = MAX[(1.35∙(0.56+25.11)+1.5∙0.7∙12.15);
(1.35∙0.85∙(0.56+25.11)+1.5∙12.15)] = 47.68 kN/m
(CC2 with ξ = 0.85, Eq. 6.10 (a) + (b) EN 1990)
Sketch
The beam can be part of a frame, only the beam will be studied. Storey height 3.6 m.

Figure 6-12. Static system

47.68 kN/m

Commentary
The beam is laterally restrained to prevent out of plane and lateral torsional buckling.
Further lateral restraints have to be provided at all plastic hinge locations. Effects of
foundation settlements, horizontal (wind) loads, normal forces and more unfavourable
combination of mobile loads are disregarded. Elastic displacements are limited till L/250 in
the rare combination and an absolute value of 10 mm for the mobile loads. Horizontal
displacements at the ends of the frame are blocked by stiff concrete cores or wind bracings.
Determination of internal forces and moments under normal conditions of use
Most practical design starts from an elastic distribution of forces by the aid of commercial
(ordinary) software tools or even analytical expressions. This explains the reason why in
the following a global elastic analysis is followed.

125
Elastic global analysis under NC
The moment and shear force redistribution of a four-span beam, equally loaded gives:

MAB = 0.077∙qd∙L²
MB = -0.107∙qd∙L²
MBC = 0.036∙qd∙L²
MC = -0.071∙qd∙L²
VBA = -0.607∙qd∙L
VBC = 0.536∙qd∙L
VCB = -0.464∙qd∙L

Reactions in ULS:

For class 1 and 2 section a limited (15%) redistribution of forces is allowed according to
clause 5.4.1.(4) from EN 1993-1-1. The internal forces and moments in the frame remain
in equilibrium with the applied loads, or:

M*AB = 0.084∙qd∙L²
M*B = -0.091∙qd∙L²
M*BC = 0.044∙qd∙L²
M*C = -0.071∙qd∙L²
V*BA = -0.591∙qd∙L
V*BC = 0.520∙qd∙L
V*CB = -0.480∙qd∙L

Reactions in ULS:

Design of cross section under NC


Steel grade S355 EN10025 or ε = 0.81, section IPE360; tf = 12.7 mm, tw = 8 mm and r =
18 mm

Flange: c/tf = ((170-8-2∙18)/2)/12.7 = 4.96 < 9ε = 7.29  class 1 (6-11)

Web: c/tw = (360-2∙12.7-2∙18)/8 = 37.33 < 72ε = 58.32  class 1 (6-12)

Shear
𝐴𝑣𝑧;𝐼𝑃𝐸360 ∙𝑓𝑦𝑘 2976∙355
VEd,max = 0.607∙47.68∙8.1 = 234.4 kN < VRd,IPE360 = = = 609.96 kN (6-13)
𝛾𝑀0 √3 1∙√3

To avoid shear buckling:

hw/tw < 72∙ε/η or (360-2∙12.7)/8 = 41.83 < 72∙0.81/1.2 = 48.6 (6-14)

 fulfilled

126
Bending moment

VEd,max < 50% VRd,IPE360 (6-15)


 no interaction with bending moment

MEd,max = 0.107∙47.68∙8.1² = 334.726 kNm and/or (6-16)

M*Ed,max = 0.091∙47.68∙8.1² = 284.674 kNm (6-17)

𝑊𝑝𝑙;𝐼𝑃𝐸360 ∙𝑓𝑦𝑘 1019∙355


MRd,IPE360 ≥ = = 361.745 kNm > MEd,max (6-18)
𝛾𝑀0 1

 both solutions are possible


Determination of internal forces and moments due to a column loss scenario
Accidental load:

qacc = 0.56+25.11+0.3∙12.15 = 29.31 kN/m (6-19)


(Eq. 6.11 EN 1990 with the use of Ψ2)
Dynamic amplification is here neglected and will be treated subsequently.
Elastic global analysis under CL without dynamic amplification
The moment and shear force redistribution of a four-span beam, equally loaded gives:

MAB = 0.024∙qacc∙L²
MB =- 0.281∙qacc∙L²
MC = 0.219∙qacc∙L²
VBA = -0.781∙qacc∙L
VBC = 1.000∙qacc∙L

Reactions in the accidental combination give for this example lower values as in the normal
ULS design situation:

With limited (15%) redistribution of forces and equilibrium with the applied loads, this
becomes:

M*AB =
0.034∙qacc∙L²
M*B =-
0.239∙qacc∙L²
M*C =
0.261∙qacc∙L²
V*BA = -
0.739∙qacc∙L
V*BC = 1.000∙qacc∙L

Reactions in the accidental combination give for this example lower values as in the normal
ULS design situation:

127
Remark; plastic global analysis means MB,pl = MC,pl = (qacc∙(2L)²/8)/2 = 0.250∙qd∙L², VBA,pl
= -0.750∙qd∙L and VBC = 1.000∙qd∙L  continue with this optimal values as the section is
class 1.
Verification of cross section under CL without dynamic amplification
Shear

VAcc,pl = 1∙29.31∙8.1 = 237.452 kN < VRd,IPE360 = 609.96 kN (6-20)

To avoid shear buckling:

hw/tw < 72∙ε/η or (360-2∙12.7)/8 = 41.83 < 72∙0.81/1.2 = 48.6 (6-21)

 fulfilled
VEd,max = 234.42 kN > 10% VRd,IPE360 = 60.996 kN and plastic global analysis  web
stiffeners should be provided within a distance along the member of h/2 from the plastic
hinge location, where h is the height of the cross section at this locations.
Bending moment

VEd,max < 50% VRd,IPE360 (6-22)


 no interaction with bending moment

MAcc,pl = 0.250∙29.31∙8.1² = 480.757 kNm (6-23)

MRd,IPE360 = 361.745 kNm < 480.757 kNm


 plastic hinges are forming which gives cause for the formation of a mechanism.
Application of the proposed tying force method can offer a solution to fulfil robustness
requirements.
Tying force requirement due to a column loss scenario via double-span beams
General
The general formulation is given by Eq. (1). The maximum equivalent force coming from
the single beam IPE360 (S355 and 7273 mm²) is based on the tensile yield strength or:

T = F = fy∙AIPE360 = 355∙7273 = 2581.92 kN (6-24)

According to Table 6-8. with equal spans the intensity factor if = 2.5 and the equivalent
load P = q∙L or P = 29.31∙8.1 = 237.411 kN

128
Table 6-8. Tying parameters for double span beams.

Without further knowledge about the connections between beams and columns a safe
assumption of α should be respected, see proposition in clause 3.6.3.2; α = 0.15 rad and
𝛼̅ = 0.15/0.20 = 0.75.
At this moment in the assessment procedure this factor η should be taken equally to 2.
The possible reduction factor ρ should be taken equally to 1 as there is no proven strain-
hardening and interaction between tying and flexural action at this moment.
Substituting the relevant values in Eq. (A.3-1) gives:

2.5
𝑇 = 2581.92 𝑘𝑁 ≥ 2 ∙ 1 ∙ ( ) ∙ 237.41 = 1582.73 𝑘𝑁 (6-25)
0.75

Resistance surrounding structure to horizontal tying


The resistance of the surrounding structure should be checked is based on the ultimate
strength of the ties subject to a corresponding material factor. Steel S355 do have a
specified yield strength of 355 N/mm² and ultimate strength of 510 N/mm². To be
complete no reduction according to the thickness of the flange or web is plied, but is
allowed.

Fu = 510∙7273 = 3709.23 kN (6-26)

The vertical distance between the effective compressive centres of rotation before
displacement = effective depth deff of the section. For an IPE section this is the distance
between the centroids of the flanges or in this example 360-12.7 = 347.3 mm

1.1.

deff
IPE360

The elastic extension δ of the beam under the axial load Fu is equal to

Fu/(EA/L) = 3709.23∙103/(210000∙7273/8100) = 19.7 mm (6-27)

129
Maximum displacement of the surrounding structure:

2
𝐿1 𝑑𝑒𝑓𝑓 𝐿1 + 𝐿2
𝑢≤ (𝛼 − ) ( )−𝛿 (6-28)
2 𝐿1 𝐿2

8100 347.3 2 16200


𝑢≤ (0.15 − ) ( ) − 19.7 = 73 𝑚𝑚 (6-29)
2 8100 8100

The concrete cores or wind bracings at the ends of the frame should limit the horizontal
displacements in the accidental load case to 73 mm to allow the development of catenary
action.
Assume at each end a concrete core (C30/37) with a second moment of area = 5.024 m4
(4x2,4 m² outer dimensions and walls of 20 cm), bending around the strong axes with at
3 levels the load F or by approximation the displacement of a cantilever beam with a height
of 3∙3.6 = 10.8 m with a distributed load of 3709.23/3.6 = 1030.3 kN/m gives 10.6 mm
horizontal displacement = ucore.
Between the cores and starting point of the catenary action there is still 8.1 m steel beam
which will be also subjected to the force F. The elongation of this part ubeam can be
calculated as before:

𝐹 3709230 ∙ 8100
𝑢𝑥,𝑏𝑒𝑎𝑚 = = = 19.7 𝑚𝑚 (6-30)
𝐸𝑠 𝐴⁄ 210000 ∙ 7273
𝐿

The total displacement ueff = ucore + ubeam = 10.6+19.7 = 30.3 mm smaller as 73 mm or


criterion fulfilled.
Minimum rotational ductility for activation of tying
The minimum rotational ductility becomes with previous values substituted:

𝑑𝑒𝑓𝑓 2(𝑢𝑒𝑓𝑓 + 𝛿) 𝐿2 347.3 2(34 + 19.7) 8100


𝛼𝑚𝑖𝑛 = +√ ( )= +√ ( ) = 0.124 𝑟𝑎𝑑 (6-31)
𝐿1 𝐿1 + 𝐿2 𝐿1 8100 16200 8100

The assumed α = 0.15 rad attends to be bigger as αmin or the condition to develop catenary
action is fulfilled.
Dynamic amplification
Validation of the conservative approach η = 2 can be made by the application of the energy
balance principles can be used as follows:
Proportion of the load resisted by flexural action λf = Pf/P = part of the load that can be
taken by flexural action, even with a plastic redistribution MRd,IPE360 = 361.745 kNm =
Pf2L/8 or Pf = 178.64 kN which leads to Pf/P = 178.64/237.41 = 0.75. Note; MRd,pl/MAcc,pl
= 361.745/480.757 = 0.75.
Proportion of the load that can be resisted by tying action under static conditions with the
chosen profile; T = 2581 kN and Pt = 2581∙0.75/(1∙2.5) = 774.57 kN. The proportion of
the load that can be resisted by tying action under static conditions will be λ t = Pt/P =
774.57/237.41 = 3.26.
Relative load ratio λ = λf/λt = 0.75/3.26 = 0.23
Verification is needed if αmin = 0.124 rad attends to be bigger or smaller as λα = 0.231∙0.15
= 0.035 rad. It is bigger (αmin > λα) so the dynamic amplification factor becomes:

130
2 2
𝜂= 2 = 2 = 2.14 ≤ 2 ?
1 𝛼 2 − 1 (0.124 − 0.231)
(6-32)
1 + 𝜆2 − ( 𝑚𝑖𝑛 − 𝜆) 1 + 0.231
3 𝛼 3 0.15

With a value higher as the theoretical maximum of 2, the assumed value of 2 will be kept.
Resistance surrounding structure to redistributed vertical gravity loads
Amplification of the gravity loading must be considered to the surrounding structure,
especially the neighbouring columns. Arising from the sudden loss of a specific column/load
bearing member. From Table 6-9 an amplification factor 0.25+0.75∙1.955 = 1.716 can be
derived.
Table 6-9 – Amplification of redistributed vertical gravity loading for double span beams.

Previously the design load of these columns in ULS was equally to (0.607+0.536)∙q d∙L =
441.434 kN (elastic analysis without redistribution). In the accidental case this became
(0.75+1.00)∙qacc∙L = 408.38 kN (plastic) in combination with the amplification this
becomes however: (0.75+(0.25+0.75∙2)∙1.00)∙qacc∙L = 583.40 kN. If the column was
optimal designed (Unity Check of 1) in ULS, it will be overloaded by a factor of 1.32! 
Additional check of the column and foundation capacity is needed.
Assessment
Global analysis
It is inevitable that the deformations will be concentrated in the joint, and hence we accept
them as the weakest link; what is important though is to know their deformation capacity.
As the capacity of the section is sufficiently large, the connections must be verified to be
able to develop a rotation of at least 0.15 rad and to resist a tying force of at least (or the
capacity of the section):

T ≥ 1582.74 kN (6-33)

The resistance of the surroundings submitted to horizontal tying was also proofed.
Details
Web stiffeners should be provided within a distance along the member of 360/2 = 180 mm
from the plastic hinge location.
Where the cross-section of the member varies along their length (i.e. openings in beams),
the following additional criteria should be satisfied:
— Adjacent to plastic hinge locations, the thickness or section of the web should not be
reduced for a distance each way along the member from the plastic hinge location of
at least 2d, where d is the clear depth of the web at the plastic hinge location see clause
5.6 of EN 1993-1-1.
— Adjacent to plastic hinge locations, the compression flange should be Class 1 for a
distance each way along the member from the plastic hinge location of not less than
the greater of:

131
o 2d, where d is as defined in a) just above.
o the distance to the adjacent point at which the moment in the member has
fallen to 0,8 times the plastic moment resistance at the point concerned.
— Elsewhere in the member the compression flange should be class 1 or class 2 and the
web should be class 1, class 2 or class 3.
It should be proofed that the connections with the columns can still resist to a tensile force
equal to T = 1582.74 kN after yielding due to bending.
Measures should be taken that no other failure mechanism (even not due to overstrength)
can cause a failure than those verified in the assessment procedure.

132
6.1.4.3 Timber Structure Example
In this example, a 4 storeys 5 × 5-bays post-and-beam timber structure is considered.
Plan and elevation views are shown in Figure 6.13 and Figure 6.14, respectively. The
structure is used as offices (building category B according to BS EN 1991-1-1, 2002) and
falls in consequence class 2a according to BS EN 1991-1-7 (2006). Span lengths along x-
and y-direction are equals to 8 and 6 m, respectively, while the inter-story height is equal
to 3.6 m. Continuous multi-storey glued laminated columns are adopted. Glulam is used
for both primary and periphery beams, labelled as B1 and B2 in Figure 6-13, respectively.
CLT slabs 295 mm thick are adopted. Red cross in Figure 6.13 and Figure 6.14 indicates
the column suddenly loss, while the dashed rectangle indicates the substructure under
consideration.

Figure 6-13. Schematic plan view of the post-and-beam timber building (adapted from Lyu et al.,
2020; dimensions in mm) (Martinelli & Izzuddin, 2022)

Source: Martinelli & Izzuddin (2022)

Figure 6-14. Schematic elevation view of the post-and-beam timber building (dimensions in mm)

Source: Martinelli & Izzuddin (2022)

133
With reference to primary beam B1, the beam self-weight, the self-weight of CLT panels,
the superimposed dead load and the live load are equal to:
— Self-weight: gk = 0.72 kN/m (assume a homogeneous glulam beam 250x720 mm of
class GL28h with 𝜌𝑘 = 410 kg/m3)
— Dead load (self-weight of CLT panels): pk = 6.0×1.45 = 8.7 kN/m (assume a 295 mm
thickness 𝜌𝑘 = 500 kg/m3);
— Superimposed Dead load: pk = 6.0×1.0 = 6.0 kN/m;
— Live load B: qk = 6.0×3 = 18.0 kN/m (office building EN 1991‐1‐1; Ψ0 = 0.70, Ψ1 =
0.50 and Ψ2 = 0.30)
Figure 6-5 shows the cross-sections of the main structural elements C1, B1 and B2. A
GL28h class was selected for the glulam structural elements having a characteristic density
𝜌𝑘 = 410 kg/m3, and values of mean and fifth percentile elastic modulus equal to 𝐸0,𝑚𝑒𝑎𝑛 =
12600 MPa 𝐸0,05 = 10200 MPa. CLT panels have a density 𝜌𝑘 = 500 kg/m3.

Figure 6-15. Cross-sections of beams and columns (dimensions in mm)

Source: Martinelli & Izzuddin (2022)

A service class 1 is assumed (EN 1995-1-1, 2004) providing a deformation factor 𝑘𝑑𝑒𝑓 = 0.6
and modification factors (𝑘𝑚𝑜𝑑 ) for permanent, medium-term and instantaneous load
duration class equals to 𝑘𝑚𝑜𝑑 = 0.6, 𝑘𝑚𝑜𝑑 = 0.8, and 𝑘𝑚𝑜𝑑 = 1.1, respectively. Material safety
factors for fundamental and accidental load combinations are set equal to 𝛾𝑀 = 1.25 and
𝛾𝑀 = 1.0, respectively. Strength values for permanent (𝑘𝑚𝑜𝑑 = 0.6; 𝛾𝑀 = 1.25), transient
(𝑘𝑚𝑜𝑑 = 0.8; 𝛾𝑀 = 1.25) and accidental (𝑘𝑚𝑜𝑑 = 1.1; 𝛾𝑀 = 1.0) design situations are listed in
Table 6-10.
Determination of internal forces and moment under normal condition of use
Most practical design starts from an elastic distribution of forces by the aid of commercial
(ordinary) software tools or even analytical expressions. Reason why in the following a
global elastic analysis is followed.

Ultimate limit state


Load combination I

The design load is calculated as (Eq. 6.10 (a) + (b) EN 1990):

1.35 × (0.72 + 8.7 + 6.0) + 1.5 × 0.7 × 18 = 39.7 kN/m


𝑞𝑑,𝐼 = 𝑚𝑎𝑥 {
0.85 × 1.35 × (0.72 + 8.7 + 6.0) + 1.5 × 18 = 44.7 kN/m
= 44.7 kN/m (6-34)

134
Table 6-10. Strength values for permanent (𝑘𝑚𝑜𝑑 = 0.6; 𝛾𝑀 = 1.25), transient (𝑘𝑚𝑜𝑑 = 0.8; 𝛾𝑀 = 1.25)
and accidental (𝑘𝑚𝑜𝑑 = 1.1; 𝛾𝑀 = 1.0) design situations

Characteristic values Design values

kmod = 0.6; kmod = 0.8; kmod = 1.1;

M = 1.25 M = 1.25 M = 1.0

fm,k (MPa) 28.0 fm,d (MPa) 13.44 17.92 30.8

ft,0,k (MPa) 19.5 ft,0,d(MPa) 9.36 12.48 21.45

ft,90,k (MPa) 0.45 ft,90,d (MPa) 0.216 0.288 0.495

fc,0,k (MPa) 26.5 fc,0,d (MPa) 12.72 16.96 29.15

fc,90,k (MPa) 3.0 fc,90,d (MPa) 1.44 1.92 3.3

fv,k (MPa) 3.2 fv,d (MPa) 1.536 2.048 3.52

Figure 6-16 shows the loading scheme for the glued laminated beam under study.

Figure 6-16. Loading scheme for the single-span beam B1 (dimensions in mm)

Source: Martinelli & Izzuddin (2022)

The shear force at the supports, the maximum bending moment at mid-span and the
reactions are reported below:
Shear forces:
1 1
𝑉𝐴𝐵 = 𝑉𝐵𝐶 = + 𝑞 ∙ 𝐿; 𝑉𝐵𝐴 = 𝑉𝐶𝐵 = − 𝑞 ∙ 𝐿;
2 2
(6-35)

Bending moments:
1
+
𝑀𝑚𝑎𝑥 = 𝑞 ∙ 𝐿2 (6-36)
8
Reactions:
1
𝑅𝐴 = 𝑅𝐶 = 𝑞 ∙ 𝐿; 𝑅𝐵 = 𝑞 ∙ 𝐿 (6-37)
2
The maximum bending moment at mid-span is equal to: 𝑀𝑑,𝐼 = 0.125 ∙ 44.7 ∙ 82 = 357.6 kNm
The shear force at the support is equal to: 𝑉𝑑,𝐼 = 0.5 ∙ 44.7 ∙ 8 = 178.8 kN
Since the load combination includes actions belonging to different duration classes, it will
be necessary to choose the value of 𝑘𝑚𝑜𝑑 which corresponds to the actions of shorter
duration; for this load combination the value for the medium-term duration must therefore
be used: 𝑘𝑚𝑜𝑑,𝐼 = 0.8
Load combination II

135
In this load combination only the dead load are considered:
𝑞𝑑,𝐼𝐼 = 1.35 × (0.72 + 8.7 + 6.0) + 0 × 18 = 20.8 kN/m (6-38)
The maximum bending moment at mid-span is equal to:
𝑀𝑑,𝐼𝐼 = 0.125 ∙ 20.8 ∙ 82 = 166.4 kNm (6-39)
The shear force at the support is equal to:
𝑉𝑑,𝐼𝐼 = 0.5 ∙ 20.8 ∙ 8 = 83.2 kN (6-40)
In this case, only permanent loads are acting, thus the value 𝑘𝑚𝑜𝑑,𝐼𝐼 = 0.6 is adopted

Verification of failure conditions of the timber beam and connectors

The most severe load combination for both the bending and shear checks is the one that
include both dead and live loads (Load combinations I). Bending and shear checks are
satisfied but omitted here for sake of brevity. Readers can refer to the Report of the Project
Team WG6.T2 for the complete checks.
The factored design shear force of the beam-to-column connectors under medium-term
actions (𝑉𝑑,𝐼 ) is 178.8 kN for the building under study.
Service limit states

The service load is equal to (Eq. 6.14 EN 1990):


q = 0.72+14.7+18.0 = 33.4 kN/m (6-41)

Deflection checks are also satisfied but omitted here for sake of brevity. The complete
derivation is reported in the Report of the Project Team WG6.T2.

Column loss scenario


The typical failure modes for which a design strategy for robustness is applicable are:
1. failure/loss of the column only (the beam-column node remains intact): in this case,
robustness exploits the coupled behaviour between:
(a) residual tensile strength of the beam-column connection which allows the activation
of a catenary behaviour of the beam itself (second order effects);
(b) membrane resistance of the floor (taking care to ensure adequate resistance of the
connections in the presence of this state of stress);
2. failure of the beam-column node: in this case, the robustness is guaranteed exclusively
by the membrane behaviour of the floor.
In this example, the loss of the column only is assumed (the beam-column node remains
intact) and the robustness relies on the catenary action of the beam (point 1a)).
The accidental load combination (Eq. 6.11 EN 1990 with the use of 2) is equal to:

𝑞𝑎𝑐𝑐 = 15.42 + 0.3 × 18 = 20.8 kN/m (6-42)

Elastic global analysis under Column Loss


Figure 6-17 shows the loading scheme for the glued laminated beam under study in case
of a central column loss (column E3 in6-14). In this case, the substructure shown in Figure
6-17 is unstable and can carry the load only in a deformed configuration.

136
Figure 6-17. Loading scheme for the beam B1 under central column loss (dimensions in mm)

Source: Martinelli & Izzuddin (2022)

The reactions without dynamic amplification due to the sudden column loss are equal to:

1
𝑅𝐴 = 𝑅𝐶 = 𝑞 ∙ 2𝐿 = 166.4 kN (6-43)
2

Application of prescriptive tying force method can offer a solution to fulfil robustness
requirements.

Tying force requirement due to a column loss scenario via double‐span beams
Equivalent load and intensity factor
Table 6-1 provides tying parameters for the double-span beam under study:
Intensity factor: 𝑖𝑓 = 2.5
Equivalent load: 𝑃 = 𝑞 ∙ 𝐿 = 20.8 ∙ 8 = 166.4 kN
Equivalent tying force: 𝑇 = 𝐹

Rotation ductility
In post-and-beam timber buildings, the lateral stability is provided by either shear core,
bracing members, or both. As a consequence, the beam-to-column connections are only
designed to carry the shear loads. Nevertheless, it does not imply that these connectors
are pinned and cannot resist an applied moment. In the case of a column loss scenario,
bending may be applied to these connections. Moreover, shear connectors are in timber
buildings mostly made by screws or nails, those components are mainly designed in
bending. With this as background, the behaviour can be in a certain way described as
ductile. Shear connectors are thus assumed in this example.
Based on the experimental data of Lyu et al. (2020), a beam rotation of 8° is assumed
that corresponds to α = 0.14 rad and 𝛼̅ = 0.14/0.20 = 0.70.

Reduction factor
CLT floor panels which could remain intact after element removal may contribute to the
load distribution after column removal. The contribution of the CLT floor panel can be
accounted in two ways:
— The first option includes directly the contribution of the CLT panels by using the general
formulation expressed in Eq. (1), where the tying via beam and the tying via floor
system are both considered and superimposed.
— The second option includes indirectly the contribution of the CLT panels by the aid of
the ρ-factor.

137
In this example the option b) has been selected and a value of 𝜌 lower than one could be
assumed. Nevertheless, given the limited knowledge of the CLT effect, a conservative
approach is followed and a value of ρ equal to 1 is chosen.

Dynamic amplification
A conservative approach is adopted in this example and a dynamic amplification factor 𝜂 =
2 has been used.

Assessment
Surrounding structure
The maximum displacement of the surrounding structure can be evaluated with Eq. (3-2)
of Report of the Project Team WG6.T2 with 𝐿1 = 𝐿2 = 𝐿:

2
𝐿 𝑑𝑒𝑓𝑓 2𝐿 (6-44)
𝑢≤ (𝛼 − ) ( )−𝛿
2 𝐿 𝐿

where 𝑑𝑒𝑓𝑓 is the vertical distance between the effective compressive centres of rotation
(pivots) for the end and internal hinges under bending action. In the example under study,
the plastic hinges are in the connections. The evaluation of 𝑑𝑒𝑓𝑓 strictly depends on the
type of connections. Assuming a shear connectors at a first approximation 𝑑𝑒𝑓𝑓 can be
evaluated as: 𝑑𝑒𝑓𝑓 = 0.8 ∙ ℎ = 576 mm

The axial force 𝐹 that produces the elastic extension 𝛿 of the beam is equal to:

𝑖𝑓 2.5 (6-45)
𝐹 = 𝜂∙𝜌∙( )∙ 𝑃 = 2∙1∙( ) ∙ 166.4 = 1188.6 kN
𝛼̄ 0.70

The elastic extension 𝛿 of the beam is equal to:

𝐹 1188.6 × 103 (6-46)


𝛿= 𝐿= 8000 = 4.2 mm
𝐸𝐴 12600 × (250 × 720)

𝐹 1188.6×103
𝛿= 𝐿= 8000 = 4.2 mm
𝐸𝐴 12600×(250×720)

The maximum displacement of the surrounding structure is equal to:

8000 576 2 16000 (6-47)


𝑢≤ (0.14 − ) ( ) − 4.2 = 32.8 mm
2 8000 8000

As general remark, a core on one side may results not sufficient; the combined axial
displacements from both end of the tie must be considered. These displacements may be
assessed considering the joint details and diaphragm action from the CLT panels. In this
example, we assume that there is enough stiffness so that the combined axial
displacements under F are less than ~33 mm.

Timber beam
The capacity of the timber beam must be verified to satisfy the following expression:

𝑖𝑓 (6-48)
𝑇≥𝜂 𝜌 ( ) 𝑃
𝛼̄

138
The maximum tensile force in the beam is equal to:

𝑇 = 𝑓𝑡,0,𝑑 ∙ 𝐴 = 21.45 ∙ (250 × 720) = 3861 kN


(6-49)

In case of a sudden column loss, the tying force requirement in the timber beam reads as:

2.5 (6-50)
𝑇 = 3861 kN ≥ 2 ∙ 1 ∙ ( ) ∙ 166.4 = 1188.6 kN → fulfilled
0.70

Timber column
Amplification of the gravity loading arising from the sudden loss of a specific column/load
bearing member must be considered to the surrounding structure, especially the
neighbouring columns. Table 3.4 of Report of the Project Team WG6.T2 provides an
amplification factor 0.25+0.75×2 = 1.75.
Previously the design load of internal column E2 (see plan view in Figure 6-13) in ULS was
equal to:

𝑁𝑑,𝐸2 = (0.5 + 0.5) ∙ 𝑞𝑑 ∙ 𝐿 = 1 ∙ 44.7 ∙ 8 = 357.6 kN (6-51)

having adopted an elastic analysis. In the accidental load condition this becomes:

𝑁𝑎𝑐𝑐,𝐸2 = 0.5 ∙ 𝑞𝑎𝑐𝑐 ∙ 𝐿 + 1.75 ∙ 1.0 ∙ 𝑞𝑎𝑐𝑐 ∙ 𝐿 = 0.5 ∙ 20.8 ∙ 8 + 1.75 ∙ 1.0 ∙ 20.8 ∙ 8 = 374.4 kN (6-52)

It is worth noting that in the above expression only the double span contribution needs to
be amplified by the factor 1.75. The column is overloaded by a factor of 1.05 thus meaning
that additional check of the column and foundation capacity are needed.

Connectors
The factored design shear force of the beam-to-column connectors under medium-term
actions was calculated previously and was equal to:

𝑅𝐴,𝑑 = 178.8 kN (6-53)

In case of column loss, the shear force at the beam-to-column connectors taking into
consideration an amplification factor of 1.75 is equal to:

1 (6-54)
𝑅𝐴,𝑎𝑐𝑐 = 1.75 ∙ 𝑞𝑎𝑐𝑐 ∙ 2𝐿 = 1.75 ∙ 0.5 ∙ 20.8 ∙ 16 = 291.2 kN
2

The connectors is overloaded in shear by a factor of 1.62.


Moreover, it should be proved that the connectors can resist, in addition to the shear force,
to a tensile force larger than:

2.5 (6-55)
𝑇𝑐𝑜𝑛𝑛 ≥ 2 ∙ 1 ∙ ( ) ∙ 166.4 = 1188.6 kN
0.70

The axial capacity of the timber beam is in general much higher than the axial capacity of
the connectors. For this reason, the axial tensile bearing capacity of the connector probably
represents the most severe check condition for the system. In a column loss scenario,
compared to an ordinary condition, there must be additional connectors to absorb the
horizontal forces. Since the column is not able to transfer tension forces perpendicular to
the fibres, a good strategy is represented by connectors that pass through the column and
then fixed into the beam.

139
The last equation can be considered a conservative approach in design the connector under
a column loss scenario for the following reasons:
— The largest dynamic amplification factor 𝜂 = 2 was assumed in the example;
— The potential tying via the CLT floor system was completely ignored; the maximum
value of the reduction factor 𝜌 = 1 was assumed in the example;
In the example under study, a beam rotation of 8° (α = 0.14 rad) was assumed based on
the data of commercially available connectors reported by Lyu et al., (2020) not specifically
designed to resist the loss of a column through catenary action.

6.2 Bridge Structures


Agarwal et al. (2012) discuss 20 structural failures, paying particular attention to structural
robustness. Faulty design and construction practices determined the collapse of an under
construction cantilever steel bridge (Quebec Bridge) in 1907 in Canada. The Almo Bridge,
a tubular steel arch bridge, built in 1960 in Sweden, collapsed in 1980 following a ship
impact. A 3-span composite concrete-steel road bridge, built in 1961 in Czech Republic,
collapsed in 2008 when under reconstruction, due to local overloading of temporary
supports. The Haeng-Ju Grand Bridge in Seoul, Korea, is another example of bridge failure
during construction. Eleven spans of the continuous pre-stressed concrete girder bridge,
collapsed in 1992 due to the failure of a temporary pier. The overall conclusion is that the
type/shape of the structure as well as the (design and construction) management practices
determine the structural robustness and by default the bridge safety. General
considerations with respect to structural robustness of bridges can be found in Starossek
(2009), advocating the consideration of design strategies that are related to segmentation
or the prevention of local failure.
Wisniewski et al. (2006) formulate a simplistic and efficient deterministic approach to
evaluate the load carrying capacity of and existing railway bridge from a robustness
perspective. The method comprises two steps: (1) analysis of an individual member that
is critical to the structural integrity of the bridge and (2) consider the system response,
following the failure of a member, to check the functionality, ultimate and damage
condition limit states in order to assess the safety of the system. Step 2 presumes to
quantify the robustness of the structure, as a function of redundancy ratio factors (rf for
serviceability; ru and rd for ultimate and damaged condition limit states). A total
redundancy factor, Φred, can then be estimated and used to assess the overall safety of the
bridge. The case study refers to the Brunna Bridge, built in 1969, a reinforced concrete
frame bridge, stretching over four spans of various lengths. The U-shape girder, skewed
at approximately 50°, supports a single railway track. The approach does not consider the
effects of the skewness and the properties of concrete and steel are considered to be time
invariant. The analysis is performed for the bridge being in two condition states:
undamaged and damaged. The redundancy of the system is found to be greater than 1 for
both condition states. This confirms the safety of the bridge, despite member failure. The
authors pointed out the high potential of this methodology if included in the maintenance
policies, expanding them from safety assessment at member level to system level.
Cavaco (2009) proposes a deterministic measure to quantify the robustness of two
reinforced concrete footbridges subjected to corrosion. Robustness was assumed to be the
attribute of a structure used to quantify the level of structural performance at various levels
of damage. The analysis was performed for two simply supported beams, 14 m long and 2
m wide. A concentrated load was applied to each at mid-span. A slab was adopted for one
structure and an I-beam for the other. The load carrying capacity was evaluated for
different levels or depths of corrosion. The results obtained from the analysis were
illustrated as functions of normalized structural performance and normalized damage. The
area under the curves was assumed to represents the robustness index.
Björnsson and Thelandersson (2010) assume a probabilistic measure to determine the
structural robustness of a bridge in case of a train collision caused by a derailment. The

140
response of a multi span post-tensioned reinforced concrete bridge to such an incident was
examined. The 172 m bridge located in Malmo, Sweden, stretches over several railway
tracks and a four-lane highway. The probability of a derailment occurrence was computed
using statistical railway accidents. The hard-impact model (EN 1991-1-7: CEN, 2006) was
employed to determine the force of impact resulting from such an event. The probability
of a support failure was computed given the train impact. The probability of deck failure
was estimated following support collapse. The results indicate that two of the three
supports investigated have a substantial risk of failure following a train impact. The findings
suggested also that the deck was not design adequately to resist the failure of one of the
supports.
Olmati et al. (2013) aim to estimate the robustness of truss bridges from the structure’s
response to impact load. This method examines the effect of an individual member failure
on the overall load carrying capacity of the structure. Therefore, a coefficient known as
member consequence factor (Cf) was introduced in the structural robustness assessment.
Values of 1 for the Cf classify the element as a primary structural element, which means
that the failure of the element could cause the total collapse of the structure. Whereas, a
value close to 0 of the Cf distinguishes the member as a secondary element of the structure
and its failure does not impact the general stability of the structure. The equation adopted
to calculate the robustness index on the basis of damage scenario, Rscenario = 100 - Cscenariof,
was updated from the equation proposed in (Nafday, 2011). High values of Cf (up to 100
%) suggests that the member collapse would almost certainly result in the entire system
collapsing. On the other hand, values of the Cf on the other side of the spectrum were
assumed to indicate a good structural robustness. The approach was tested on small
theoretical structures as well as on an existing bridge. The investigation focused on the I-
35W Minneapolis steel truss bridge which was built in the early 1960s and collapsed in
2007. Different members of one of the two trusses were removed from the 2D model to
simulate damage. The structural robustness was then measured from the Cf for two cases:
one in its original form of the truss and the second for an improved form of the truss. The
same damage scenarios were considered in both situations. A low robustness index was
found from investigating the original structure.
Biondini and Frangopol (2015) employ deterministic and probabilistic approaches to
evaluate the life-cycle robustness of a reinforced concrete bridge pier with a box cross-
section. The actual damage state of the pier and performance level at different time
instants over the structural lifetime were considered in the evaluation. To conform with the
principle that the robustness index is the difference between an undamaged and damaged
system as per (Frangopol and Curley, 1987, Biondini and Frangopol, 2008), a robustness
factor was introduced. This factor was defined as a function of the performance index and
the related damage index. It was assumed that the condition of structural robustness was
met for robustness factor greater than or equal to 1; otherwise, the system is weak, if the
robustness factor is less than 1. The importance of the structure (temporary, ordinary or
strategic) is considered to impact the robustness measure and it is integrated in the
approach through the importance factor. The outcomes of this analysis showed that the
probabilistic approach validated the deterministic findings, but emphasised that “the effects
of uncertainty tend to increase over time periods when the susceptibility to damage
increases and robustness decreases”. Moreover, the evaluation of life-cycle robustness of
existing structures has the potential to give details on the deterioration impact on the
overall performance of a structure, considering the damage state and performance level at
the time of the assessment.
Moreira et al. (2016) adopt deterministic measures (such as load capacity and extent of
damage) to quantify the robustness index of a masonry arch railway bridge. They adopted
the robustness approach proposed in (Cavaco, 2009, Cavaco et al., 2013) to evaluate the
ultimate load carrying capacity of the Calharda Viaduct, built in 1882 in Portugal. Damage
scenarios such as longitudinal cracking, transversal cracking, spalled masonry arch
voussoirs and masonry deterioration and fatigue were considered in the analysis. For
different damage scenario with various percentage of damage – 0%, 10%, 25%, 50% and

141
100% damage – the robustness index was found to be close to 1, which identifies a robust
bridge.
Guimarães et al. (2017) suggest a reliability-based robustness assessment model for
bridge safety evaluation. Reliability (Pβ), damage tolerance (PD), redundancy (PR) and
ductility (PΦ) were the robustness indicators evaluated in their research in order to
determine the structural performance. The model was validated against a reinforced
concrete clamped beam. Different percentages of corrosion (25% and 40 %) were
introduced to define the damage scenarios considered. For each damage scenario, the
values of the performance indicators were obtained and displayed in a quadratic diagram,
with side length of 1, as shown in Figure 6-18. The quadrilateral surface area denoted by
the performance indicators was the normalized robustness index. It was observed that
each indicator may be quantified for its effect. The robustness index was found to be
decreased with the decrease of ductility and redundancy.

Figure 6-18. Performance indicators

Source: Guimarães et al., 2017

Finally, various authors consider the important case of loss of support of bridge structures.
Shoa et al. (2021) focus on the case of cable breaks in cable supported structures, Olmati
et al. (2012) consider extreme loads resulting in the loss of supporting elements in steel
bridge structures whilst Björnsson and Thelandersson (2010) consider robustness
evaluation of a multi-span concrete bridge crossing multiple rail tracks in the case of train
derailment at the supports. Clearly an important scenario to consider in the analysis of
bridge structure robustness the loss of support, its probability and associated
consequences should be carefully considered in robustness evaluation of bridge structures.

References
Abbasnia, R., & Nav, F. M. 2016. A theoretical method for calculating the compressive arch
capacity of RC beams against progressive collapse. Structural Concrete, 17(1), 21–31
Agarwal, J., Haberland, M., Holicky, M., Sykora, M. & Thelandersson, S. 2012. Robustness
of structures: Lessons from failures. Structural Engineering International, 22, 105-111.
Belletti B., Franceschini L., Ravasini S., (2019), Tie Force Method for Reinforced Concrete
Structures. Proceedings of the International fib Symposium on Conceptual Design of
Structures 2019: 57-64; Madrid, Spain; 26 - 28 September 2019.
Belletti B., Franceschini L., Ravasini S., (2021), Experimental Database and Validation of
Analytical Models for Tie Force and Alternate Load Path Methods for RC Structures.
CEN/TC250/WG6 Project Team WG6.T2.
Biondini, F. & Frangopol, D. 2008. Life-Cycle Civil Engineering: Proceedings of the
International Symposium on Life-Cycle Civil Engineering, IALCCE'08, held in Varenna, Lake
Como, Italy on June 11-14, 2008, CRC Press.

142
Biondini, F. & Frangopol, D. 2015. Life-cycle robustness of deteriorating concrete
structures.
Björnsson, I. & Thelandersson, S. Robustness analysis of bridge when exposed to train
collision due to derailment. COST Action C26 Final Conf, 2010. 603-608.
BS EN 1991-1-1. (2002). Eurocode 1: Actions on structures — Part 1-1: General actions
— Densities, self-weight, imposed loads for buildings. Brussels.
BS EN 1991-1-7. (2006). Eurocode 1 - Actions on structures – Part 1-7: General actions –
Accidental actions. Brussels.
Cavaco, E., Casas, J. R. and Neves, L. Quantifying redundancy and robustness of
structures. IABSE Symposium Report, 2013. International Association for Bridge and
Structural Engineering, 78-99.
Cavaco, E. S. 2009. Robustness of corroded reinforced concrete structures.
Department of Defense (2009), Unified Facilities Criteria, Design of Buildings to Resist
Progressive Collapse, UFC 4-023-03, Change 3, November 2016, Washington, DC, USA.
Dinu, F., Marginean, I., & Dubina, D. (2017). Experimental testing and numerical modelling
of steel moment-frame connections under column loss. Engineering structures, 864-878.
CEN, (2004), EN 19955: Design of timber structures – Part 1-1: General – Common rules
and rules for buildings. European Committee for Standardization (CEN): Brussels
(Belgium).
CEN, (2005), EN 1990:2002 + A1: Basis of structural design. European Committee for
Standardization (CEN): Brussels (Belgium).
CEN, (2007), EN 1999-1-1:Design of aluminium structures - part 1-1: General structural
rules. European Committee for Standardization (CEN): Brussels (Belgium).
CEN, (2006), EN 1991-1-7: Eurocode 1 - Actions on structures - Part 1-7: General actions
- Accidental actions, European Committee for Standardization (CEN): Brussels (Belgium).
Demonceau, J-F, Golea T, Jaspart J-P, Elghazouli A, Khalil Z, Santiago A, Santos A F,
Simoes da Silva, L, Kuhlmann U, Skarmoutsos G, Baldassino N, Zandonini R, Zordan M,
Dinu F, Margniean I, Jakab, D, Dubina, D, Wertz, F, Weynand, K, Obiala, R and Candeias,
M, (2022). Mitigation of the risk of progressive collapse in steel and composite building
frames under exceptional events – design manual. FAILNOMORE RFCS project, Edited by
the European Convention for Constructional Steelwork.
Demonceau, J-F and Jaspart, J-P (2010). Experimental test simulating a column loss in a
composite frame. Advanced Steel Construction Journal, Vol. 6, No. 3, pp. 891-913.
Dinu, F., Marginean, I. & Dubina, D., 2017. Experimental testing and numerical modelling
of steel moment-frame connections under column loss. Engineering structures, pp. 864-
878.
EUR 28906. (2017). Standardisation of safety assessment procedures across brittle to
ductile failure modes. Luxembourg: European commission, Directorate-General for
Research and Innovation.
Fang, C., Izzuddin, B.A., Elghazouli, A.Y., and Nethercot, D.A., (2013), "Modeling of Semi-
Rigid Beam-to-Column Steel Joints under Extreme Loading", Frontiers of Structural and
Civil Engineering, Vol. 7, No. 3, pp. 245-263.
Farazman, S., Izzuddin, B.A., and Cormie, D., (2013), "Influence of Unreinforced Masonry
Infill Panels on the Robustness of Multistory Buildings", Journal of Performance of
Constructed. Facilities, ASCE, Vol. 27, No. 6, 2013, pp. 673-682.
Favvata, M., Izzuddin, B.A., and Karayannis, C.G., (2008), "Modelling Exterior Beam-
Column Joints for Seismic Analysis of RC Frame Structures", Earthquake Engineering and
Structural Dynamics, Vol. 37, No. 13, pp. 1527-1548.

143
Frangopol, D. M. and Curley, J. P. 1987. Effects of damage and redundancy on structural
reliability. Journal of structural engineering, 113, 1533-1549.
Gudmundsson, G.V., and Izzuddin, B.A., (2010), "The ‘Sudden Column Loss’ Idealisation
for Disproportionate Collapse Assessment", The Structural Engineer, Vol. 88, No. 6, 2010,
pp. 22-26.
Guimarães, H., Fernandes, J., Matos, J. and Henriques, A. Robustness assessment—A new
perspective to achieve a performance indicator. (2017). 14th International Probabilistic
Workshop. Springer, 499-507.
Izzuddin, B.A., (2010), "Robustness by Design – Simplified Progressive Collapse
Assessment of Building Structures", Stahlbau, Vol. 79, No. 8, pp. 556-564.
Izzuddin, B.A., (2012), "Mitigation of Progressive Collapse in Multi-Storey Buidlings",
Advances in Structural Engineering, Vol. 15, No. 9, pp. 1505-1520.
Izzuddin, B.A. (2022), “Rational Robustness Design of Multistory Building Structures”,
Journal of Structural Engineering, ASCE, 148(3), 04021279.
Izzuddin, B.A., and Elnashai, A.S., (1993), "Adaptive Space Frame Analysis - Part II: A
Distributed Plasticity Approach", Proceedings of the Institution of Civil Engineers,
Structures and Buildings, Vol. 99, pp. 317-326.
Izzuddin, B.A., and Lloyd Smith, D., (2000), "Efficient Nonlinear Analysis of Elasto-Plastic
3D R/C Frames Using Adaptive Techniques", Computers & Structures, Vol. 78, No. 4, pp.
549-573.
Izzuddin, B.A., and Sio, J., (2022), “Rational Horizontal Tying Force Method for Practical
Robustness Design of Building Structures”, Engineering Structures, 252(1), 113676
Izzuddin, B.A., and Sio, J., (2020b), “Prescriptive Tying Force Requirements and Alternate
Load Path Methods for Building Structures”, CEN/TC250/WG6 Project Team WG6.T2.
Izzuddin, B.A., Tao, X.Y., and Elghazouli, A.Y., (2004), "Realistic Modelling of Composite
and R/C Floor Slabs under Extreme Loading – Part I: Analytical Method", Journal of
Structural Engineering, ASCE, Vol. 130, No. 12, pp. 1972–1984.
Izzuddin, B.A., Vlassis, A.G., Elghazouli, A.Y., and Nethercot, D.A., (2008), "Progressive
Collapse of Multi-Storey Buildings due to Sudden Column Loss — Part I: Simplified
Assessment Framework", Engineering Structures, Vol. 30, No. 5, pp. 1308-1318.
Knoll, F., & Vogel, T. (2009). Design for robustness. Zurich: IABSE-AIPC-IVBH.
Lyu, C. H., Gilbert, B. P., Guan, H., Underhill, I. D., Gunalan, S., Karampour, H., & Masaeli,
M. (2020). Experimental collapse response of post-and-beam mass timber frames under a
quasi-static column removal scenario. Engineering Structures, 213(March), 110562.
Martinelli, P., (2021), Considerations on the rotational ductility in timber post-and-beam
structures. CEN/TC250/WG6 Project Team WG6.T2.
Martinelli, P., Izzuddin, B. A., (2022) Validation and application of rational tying method
for robustness design of post-and-beam timber buildings, Wood Material Science &
Engineering
Masaeli, M., Gilbert, B. P., Karampour, H., Underhill, I. D., Lyu, C. H., & Gunalan, S. (2020).
Scaling effect on the moment and shear responses of three types of beam-to-column
connectors used in mass timber buildings. Engineering Structures, 208(February), 110329.
Minga, E., Macorini, L., Izzuddin, B.A., and Calio, I., (2020), "3D Macroelement Approach
for Nonlinear FE Analysis of URM Components Subjected to In-plane and Out-of-plane
Cyclic Loading", Engineering Structures, Vol. 220, 110951.
Molkens, T. (2021). Lumped plasticity approach for the robustness assessment of
framework structures. In preparation.

144
Moreira, V. N., Fernandes, J., Matos, J. C. and Olivera, D. V. Robustness as Performance
Indicator for masonry arch bridges. COST Action TU1406: eBook of the 1st Workshop
Meeting. Geneva: ETH-Zürich, 2016.
Nafday, A. M. 2011. Consequence-based structural design approach for black swan events.
Structural Safety, 33, 108-114.
Olmati, P., Gkoumas, K., Brando, F. and Cao, L. 2013. Consequence-based robustness
assessment of a steel truss bridge. Steel and Composite Structures, 14, 379-395.
Ravasini S., Belletti B., Brunesi E., Nascimbene R., Parisi F., (2021), Nonlinear Dynamic
Response of a Precast Concrete Building to Sudden Column Removal. Applied Sciences 11,
599
Ravasini S., Sio J., Franceschini L., Izzuddin B. A., Belletti B., (2021), Validation of
simplified tying force method for robustness assessment of RC framed structures.
Engineering Structures, 249, 113291..
Report of Project Team WG6.T2 Robustness Rules in Material Related Eurocode Parts.
Mandate M/515 Development of 2nd Generation of EN Eurocodes
SCI. (2017). Design Manual for Structural Stainless Steel, 4th edition. Berkshire, UK: SCI.
Shao G., Jin H., Jiang R., Xu Y., (2021), 'Dynamic Response and Robustness Evaluation of
Cable Supported Arch Bridges Subjected to Cable Breaking', Shock and Vibration.
Steenhuis, M., Jaspart, J.-P., Gomes, F. and Leino, F. (1998), "Application of the
Component Method to Steel Joints", Proceedings of the COST C1 International Conference
in Liège, pp. 125–143, Report EUR 18854 EN. Luxembourg: Office for Official Publications
of the European Communities, 1999.
UFC 4-023-03. (2016). Design of buildings to resist progressive collapse. USA:
Departement of Defence.
Wisniewski, D., Casas, J. R. and Ghosn, M. 2006. Load capacity evaluation of existing
railway bridges based on robustness quantification. Structural Engineering International,
16, 161-166.
Xavier, F.B., Macorini, L., Izzuddin, B.A., (2015), "Robustness of Multistory Buildings with
Masonry Infill", Journal of Performance of Constructed. Facilities, ASCE, Vol. 29, No. 5,
B4014004.
Yun, X., & Gardner, L. (2017). Stress-strain curves for hot-rolled steels. Journal of
constructional steel research, 36-46.
Yun, X., Zhongxing, W., & Gardner, L. (2021). Full-range stress-strain curves for aluminum
alloys. Journal of structural eningeering, 04021060.
Zolghadr Jahromi, H., Vlassis, A.G., and Izzuddin, B.A., (2013), "Modelling Approaches for
Robustness Assessment of Multi-Storey Steel-Composite Buildings", Engineering
Structures, Vol. 51, pp. 278-294.

145
List of figures
Figure 2-1. Collapse of Ronan Point Building, London, UK (1968) ..............................14
Figure 2-2. Collapse of the Alfred Murrah Federal Building, Oklahoma City, USA (1995)
..........................................................................................................................14
Figure 2-3. Terrorist attack - The Pentagon Building, Washington D.C., USA (2001) ....15
Figure 2-4. Collapse of Terminal 2E of the Charles de Gaulle Airport, Paris, France
(2004) ................................................................................................................15
Figure 2-5. Collapse of (a) Heang-Ju Bridge, Seoul, South Korea (1992) and (b) and
Tasman Bridge, Hobart, Australia (1975) .................................................................15
Figure 2-6. Classification of threats .......................................................................22
Figure 2-7. Damage propagation mechanisms: (a) Directionality-based; (b) Adjacency-
based. .................................................................................................................25
Figure 2-8. Structural robustness and vulnerability .................................................27
Figure 2-9. System definition for a roadway network ...............................................27
Figure 3-1. Timeline of the main progressive collapse events and the developments of
design provisions. .................................................................................................36
Figure 3-2. Distinction strategies for identified and unidentified accidental actions in EN
1991-1-7 .............................................................................................................39
Figure 3-3. Recommended limit of admissible damage according to EN 1991-1-7........43
Figure 3-4. Illustration of the steps in a risk analysis for structures subject to accidental
actions according to EN 1991-1-7; (a) hazard, (b) damage, (c) collapse. .....................44
Figure 3-5. Tie forces in a frame structure. ............................................................49
Figure 3-6. Justification for the prescriptive tie rules for internal horizontal ties of
framed structures according to EN 1991-1-7 ............................................................54
Figure 3-7. Vertical equilibrium for the deformed state to derive the prescriptive tie
rules for internal horizontal ties of framed structures according to EN 1991-1-7 ...........55
Figure 3-8. Alternative Path method and strategy for the notional removal of elements
in the structure. ...................................................................................................56
Figure 5-1. Force f=f0=f1 versus displacement s of a truss system in the intact state
and after elimination of one member. (a) Force-displacement diagrams. (b) Backward
pseudo-loads (effects of repair). (c) Forward pseudo-loads (effects of damage). ..........83
Figure 5-2. Types of failure (adapted from Biondini & Restelli 2008): (a) local and (b)
global collapse. ....................................................................................................85
Figure 5-3. Overview of structural analysis methods for progressive collapse analysis
and robustness quantification. ................................................................................88
Figure 5-4. Robustness factor 𝑅 = 𝑅(𝜌, Δ). (a) Performance 𝜌 vs damage Δ state diagram
(𝛼 = 1); (b) Role of the shape parameter 𝛼 on the robustness threshold 𝑅 = 1. .............97
Figure 5-5. Comparison of time-variant robustness over a structural lifetime T for two
damage scenarios (indicators at equal time intervals t=T/5): (a) performance functions
 (), and (b) robustness factor profiles 𝑅 = 𝑅𝑡 for 𝛼 = 1. ........................................98
Figure 6-1. Development of tying and flexural resistance ....................................... 107
Figure 6-2. Experimental and analytical tie resistance at the catenary stage evaluated at
the assumed ductility level .................................................................................. 109
Figure 6-3. Engineering stress-strain relationship for steel (Yun & Gardner, 2017) .... 111

146
Figure 6-4. Engineering Typical engineering stress-strain curve for aluminium alloys . 115
Figure 6-5. Multi-storey building subject to sudden column loss .............................. 116
Figure 6-6. Sub-structural levels for robustness assessment .................................. 117
Figure 6-7. Nonlinear static pushdown response ................................................... 118
Figure 6-8. Simplified dynamic assessment and pseudo-static response (Izzuddin et al.,
2008)................................................................................................................ 119
Figure 6-9. Performance indicators ...................................................................... 120
Figure 6-10. Plan view of the emulative precast RC frame under interior column loss 121
Figure 6-11. Lateral stiffness provided by the topping slab of 5 cm ......................... 123
Figure 6-12. Static system ................................................................................. 125
Figure 6-13. Schematic plan view of the post-and-beam timber building (dimensions in
mm) ................................................................................................................. 133
Figure 6-14. Schematic elevation view of the post-and-beam timber building
(dimensions in mm) ............................................................................................ 133
Figure 6-15. Cross-sections of beams and columns (dimensions in mm) .................. 134
Figure 6-16. Loading scheme for the single-span beam B1 (dimensions in mm) ....... 135
Figure 6-17. Loading scheme for the beam B1 under central column loss (dimensions in
mm) ................................................................................................................. 137
Figure 6-18. Performance indicators .................................................................... 142

147
List of tables
Table 2-1. Definitions related to the concept of structural robustness ........................16
Table 2-2. Definitions related to the concept of disproportionate collapse ...................18
Table 2-3. Definitions related to the concept of progressive collapse ..........................18
Table 2-4. Possible hazardous events ....................................................................22
Table 2-5. Classification of failure consequences .....................................................29
Table 3-1. Consequence Classes according to EN 1990. ...........................................40
Table 3-2. Design tie forces according to EN 1991-1-7 (CEN, 2006). .........................41
Table 3-3. Design values of ties according to EN 1992-1-1 (CEN, 2005). ....................45
Table 3-4. Occupancy Categories and Design Requirements. ....................................49
Table 3-5. Summary design tie forces according to UFC 4-023-03 .............................50
Table 3-6. Load combination for alternative load path analysis according to UFC 4-023-
03. .....................................................................................................................51
Table 4-1. Classification of identified accidental actions and accidental actions associated
with unidentified hazardous events. ........................................................................63
Table 4-2. Robustness design approaches. .............................................................65
Table 5-1. Some advantages and drawbacks of different methods of structural response
analysis ...............................................................................................................92
Table 6-1. Tying parameters and redistributed load amplification for selected 1D/2D
systems ............................................................................................................ 108
Table 6-2. Recommended value of γrm .................................................................. 112
Table 6-3. Ultimate rotations for connectors Types 1-3 from bending tests at full-scale
and at ¼ scale ................................................................................................... 114
Table 6-4. Ultimate rotations for connectors Types 1-4 from push-down tests at ¼ scale
........................................................................................................................ 114
Table 6-5. Applied loads ..................................................................................... 121
Table 6-6. Tying parameters for double-span beams ............................................. 121
Table 6-7. Amplification of redistributed vertical gravity loading for double-span beams
........................................................................................................................ 124
Table 6-8. Tying parameters for double span beams. ............................................. 129
Table 6-9 – Amplification of redistributed vertical gravity loading for double span beams.
........................................................................................................................ 131
Table 6-10. Strength values for permanent (𝑘𝑚𝑜𝑑 = 0.6; 𝛾𝑀 = 1.25), transient (𝑘𝑚𝑜𝑑 =
0.8; 𝛾𝑀 = 1.25) and accidental (𝑘𝑚𝑜𝑑 = 1.1; 𝛾𝑀 = 1.0) design situations ........................ 135

148
Getting in touch with the EU
In person
All over the European Union there are hundreds of Europe Direct centres. You can find the address of the
centre nearest you online (european-union.europa.eu/contact-eu/meet-us_en).
On the phone or in writing
Europe Direct is a service that answers your questions about the European Union. You can contact this
service:
— by freephone: 00 800 6 7 8 9 10 11 (certain operators may charge for these calls),
— at the following standard number: +32 22999696,
— via the following form: european-union.europa.eu/contact-eu/write-us_en.

Finding information about the EU


Online
Information about the European Union in all the official languages of the EU is available on the Europa
website (european-union.europa.eu).
EU publications
You can view or order EU publications at op.europa.eu/en/publications. Multiple copies of free publications
can be obtained by contacting Europe Direct or your local documentation centre (european-
union.europa.eu/contact-eu/meet-us_en).
EU law and related documents
For access to legal information from the EU, including all EU law since 1951 in all the official language
versions, go to EUR-Lex (eur-lex.europa.eu).
EU open data
The portal data.europa.eu provides access to open datasets from the EU institutions, bodies and agencies.
These can be downloaded and reused for free, for both commercial and non-commercial purposes. The
portal also provides access to a wealth of datasets from European countries.
The Joint Research Centre (JRC) provides
independent, evidence-based knowledge and
science, supporting EU policies to positively
impact society

EU Science Hub
Joint-research-centre.ec.europa.eu

You might also like