You are on page 1of 248

DESIGNERS GUIDES TO THE EUROCODES

DESIGNERS GUIDE TO EUROCODE 8:


DESIGN OF BRIDGES FOR EARTHQUAKE
RESISTANCE
EN 1998-2

BASIL KOLIAS
DENCO S.A, Greece

MICHAEL N. FARDIS
University of Patras, Greece

ALAIN PECKER
Geodynamique et Structure, France

Series editor
Haig Gulvanessian CBE

Published by ICE Publishing, 40 Marsh Wall, London E14 9TP

Full details of ICE Publishing sales representatives and distributors can be found at:
www.icevirtuallibrary.com/info/printbooksales

Eurocodes Expert
Structural Eurocodes offer the opportunity of harmonised design standards for the European construction
market and the rest of the world. To achieve this, the construction industry needs to become acquainted
with the Eurocodes so that the maximum advantage can be taken of these opportunities.
Eurocodes Expert is an ICE and Thomas Telford initiative set up to assist in creating a greater awareness
of the impact and implementation of the Eurocodes within the UK construction industry.
Eurocodes Expert provides a range of products and services to aid and support the transition to Eurocodes.
For comprehensive and useful information on the adoption of the Eurocodes and their implementation
process please visit our website or email eurocodes@thomastelford.com

www.icevirtuallibrary.com
A catalogue record for this book is available from the British Library
ISBN 978-0-7277-5735-7
# Thomas Telford Limited 2012
ICE Publishing is a division of Thomas Telford Ltd, a wholly-owned subsidiary of the Institution of Civil
Engineers (ICE).
All rights, including translation, reserved. Except as permitted by the Copyright, Designs and Patents Act
1988, no part of this publication may be reproduced, stored in a retrieval system or transmitted in any form
or by any means, electronic, mechanical, photocopying or otherwise, without the prior written permission of
the Publishing Director, ICE Publishing, 40 Marsh Wall, London E14 9TP.
This book is published on the understanding that the authors are solely responsible for the statements made
and opinions expressed in it and that its publication does not necessarily imply that such statements and/or
opinions are or reect the views or opinions of the publishers. While every effort has been made to ensure
that the statements made and the opinions expressed in this publication provide a safe and accurate guide,
no liability or responsibility can be accepted in this respect by the authors or publishers.
Associate Commissioning Editor: Jennifer Barratt
Production Editor: Imran Mirza
Market Specialist: Catherine de Gatacre

Typeset by Academic Technical, Bristol


Index created by Indexing Specialists (UK) Ltd, Hove, East Sussex
Printed and bound by CPI Group (UK) Ltd, Croydon, CR0 4YY

Preface
Aim of the Designers Guide
This Designers Guide to EN 1998-2:2005 covers the rules for the seismic design of bridges,
following in a loose way the contents of this EN Eurocode. It highlights its important points
without repeating them, providing comments and explanations for its application, as well as
background information and worked-out examples. However, it does not elaborate every
single clause in EN 1998-2:2005, neither does it follow strictly the sequence of its clauses.

Layout of this guide


All cross-references in this guide to sections, clauses, subclauses, paragraphs, annexes, gures,
tables and expressions of EN 1998-2 and EN 1998-5 are in italic type, which is also used
where text from EN 1998-2 and EN 1998-5 has been directly reproduced (conversely, quotations
from other sources, including other Eurocodes, and cross-references to sections, etc., of this
guide, are in roman type). Numbers within square brackets after cross-references in the
margin refer to Parts 1, 2 and 5 of EN 1998: EN 1998-1 [1], EN 1998-2 [2], EN 1998-5 [3]. Expression numbers specic to this guide are prexed by D (for Designers Guide), for example,
Eq. (D3.1), to prevent confusion with expression numbers from EN 1998.

Acknowledgements
This Designers Guide would not have been possible without the successful completion of
EN 1998-2:2005. Those involved in the process were:
g
g

national delegates and national technical contacts to Subcommittee 8 of CEN/TC250


the Project Team of CEN/TC250/SC8 that worked for the conversion from the ENV to
the EN: namely PT4, convened by Alex Plakas.

Contents
Preface
Aim of the Designers guide
Layout of this guide
Acknowledgements

v
v
v
v

Chapter 1

Introduction and scope


1.1. Introduction
1.2. Scope of Eurocode 8
1.3. Scope of Eurocode 8 Part 2
1.4. Use of Eurocode 8 Part 2 with the other Eurocodes
1.5. Additional European standards to be used with EN 1998-2:2005
1.6. Assumptions
1.7. Distinction between principles and application rules
1.8. Terms and denitions symbols
References

1
1
1
2
2
3
4
4
4
4

Chapter 2

Performance requirements and compliance criteria


2.1. Performance-based seismic design of bridges
2.2. Performance requirements for new bridges in Eurocode 8
2.3. Compliance criteria for the non-collapse requirement and
implementation
2.4. Exemption from the application of Eurocode 8
References

5
5
7
8
16
16

Chapter 3

Seismic actions and geotechnical aspects


3.1. Design seismic actions
3.2. Siting and foundation soils
3.3. Soil properties and parameters
3.4. Liquefaction, lateral spreading and related phenomena
References

19
19
29
30
32
36

Chapter 4

Conceptual design of bridges for earthquake resistance


4.1. Introduction
4.2. General rules for the conceptual design of earthquake-resistant bridges
4.3. The choice of connection between the piers and the deck
4.4. The piers
4.5. The abutments and their connection with the deck
4.6. The foundations
References

37
37
38
43
53
59
64
65

Chapter 5

Modelling and analysis of bridges for seismic design


5.1. Introduction: methods of analysis in Eurocode 8
5.2. The three components of the seismic action in the analysis
5.3. Design spectrum for elastic analysis
5.4. Behaviour factors for the analysis
5.5. Modal response spectrum analysis
5.6. Fundamental mode analysis (or equivalent static analysis)
5.7. Torsional effects in linear analysis
5.8. Effective stiffness for the analysis
5.9. Calculation of seismic displacement demands through linear analysis
5.10. Nonlinear analysis
References

67
67
68
69
69
73
92
98
100
107
110
117
vii

Chapter 6

viii

Verication and detailing of bridge components for earthquake resistance


6.1. Introduction
6.2. Combination of gravity and other actions with the design seismic
action
6.3. Verication procedure in design for ductility using linear analysis
6.4. Capacity design of regions other than exural plastic hinges in
bridges of ductile behaviour
6.5. Overview of detailing and design rules for bridges with ductile or
limited ductile behaviour
6.6. Verication and detailing of joints between ductile pier columns and
the deck or a foundation element
6.7. Verications in the context of design for ductility based on nonlinear
analysis
6.8. Overlap and clearance lengths at movable joints
6.9. Seismic links
6.10. Dimensioning of bearings
6.11. Verication of abutments
6.12. Verication of the foundation
6.13. Liquefaction and lateral spreading
References

119
119

Chapter 7

Bridges with seismic isolation


7.1. Introduction
7.2. Objective, means, performance requirements and conceptual design
7.3. Design seismic action
7.4. Behaviour families of the most common isolators
7.5. Analysis methods
7.6. Lateral restoring capability
References

171
171
171
174
174
185
191
191

Chapter 8

Seismic design examples


8.1. Introduction
8.2. Example of a bridge with ductile piers
8.3. Example of a bridge with limited ductile piers
8.4. Example of seismic isolation
References

193
193
193
210
221
249

Index

251

119
122
124
129
129
132
135
140
142
155
159
164
167

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.001

Chapter 1

Introduction and scope


1.1.

Introduction

Design of structures for earthquake resistance penetrated engineering practice for buildings
much earlier than for bridges. There are several reasons for this. First, seismic design is of relevance mainly for piers, but is secondary for the deck. The deck, though, receives in general
far more attention than the piers, as it is more important for the function and the overall cost
of the bridge, while its engineering is also more challenging. So, seismic considerations, being
relevant mainly for the less important components of bridges, have traditionally been of lower
priority. Second, a good number of bridges are not so sensitive to earthquakes: the long-span
ones which are also the subject of lots of attention and of major design and engineering
effort are very exible, and their long periods of vibration are outside the frequency range of
usual ground motions. At the other extreme, short bridges, with one or only few spans, often
follow the ground motion with little distress, and normally suffer only minor damage.
However, with the very rapid expansion of transportation networks, the new priorities in land
use especially in urban areas and the sensitivities of recent times to protection of the environment, bridge engineering has spread from the traditional eld of short crossings of rivers, ravines
or other natural barriers or of over- and underpasses for motorways to long viaducts consisting
of a large number of spans on equally numerous piers, often crossing territories with different
ground or soil conditions. The heavy damage suffered by such types of engineering works in
the earthquakes of Loma Prieta in 1989 and Kobe in 1995 demonstrated their seismic vulnerability. More recent events have conrmed the importance of proper seismic design (or lack of
it) for bridge projects.
Owing to these developments, recent decades have seen major advances in the seismic engineering
of bridges. It may now be claimed with a certain amount of condence that the state-of-the-art in
the seismic design of bridges is catching up with that of buildings, which is more deeply rooted in
common design practice and codes. Europe, where even the moderate-to-high seismicity
countries of the south lacked modern seismic design codes for bridges, has seen the development
of EN 1998-2:2005 as a modern and complete seismic design standard, on par with its counterparts in California, Japan and New Zealand. Part 2 of Eurocode 8 (CEN, 2005a) is quite
advanced from the point of view of the state-of-the-art and of seismic protection technology,
not only compared with the pre-existing status at national levels but also with respect to the
other parts of the new European seismic design standard (EN Eurocode 8) that address other
types of civil engineering works. It is up to the European community of seismic design
practice to make good use of it, to the benet of the seismic protection of new bridges in
Europe and of its own professional competiveness in other seismic parts of the world. This
Designers Guide aspires to help this community become familiar with Part 2 of Eurocode 8,
get the most out of it and apply it in a cost-effective way.

1.2.

Scope of Eurocode 8

Eurocode 8 covers the design and construction of earthquake-resistant buildings and other
civil engineering works including bridges, but excluding nuclear power plants, offshore structures and large dams. Its stated aim is to protect human life and property in the event of an
earthquake and to ensure that structures that are important for civil protection remain
operational.

Clauses 1.1.1(1),
1.1.1(2) [1]
Clause 1.1.1(1) [2]

Eurocode 8 has six Parts, listed in Table 1.1. Among them, only Part 2 (CEN, 2005a) is covered in
this Designers Guide.

Clauses 1.1.1(4),
1.1.3(1) [1]
1

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 1.1. Eurocode 8 parts


Part

EN

Title

EN 1998-1:2004

2
3

EN 1998-2:2005
EN 1998-3:2005

4
5

EN 1998-4:2006
EN 1998-5:2004

EN 1998-6:2005

Design of structures for earthquake resistance.


seismic actions, rules for buildings
Design of structures for earthquake resistance.
Design of structures for earthquake resistance.
retrotting of buildings
Design of structures for earthquake resistance.
Design of structures for earthquake resistance.
retaining structures, geotechnical aspects
Design of structures for earthquake resistance.

1.3.
Clauses 1.1.1(2)
1.1.1(4), 1.1.1(6) [2]

General rules,
Bridges
Assessment and
Silos, tanks, pipelines
Foundations,
Towers, masts, chimneys

Scope of Eurocode 8 Part 2

Part 2 of Eurocode 8 (CEN, 2005a) has as its sole object the seismic design of new bridges. It
focuses on bridges having a deck superstructure supported directly on vertical or nearly
vertical concrete or steel piers and abutments. The seismic design of cable-stayed or arched
bridges is only partly covered, while that of suspension bridges, timber bridges (strictly
speaking, bridges on timber piers), masonry bridges, moveable bridges or oating bridges is
not covered at all. Part 2 of Eurocode 8 also covers the design of bridges with seismic isolation.
Unlike existing buildings, whose seismic assessment and retrotting is covered in Eurocode 8
(CEN, 2005b), existing bridges are not addressed at all.

1.4.
Clauses 1.1.2(1)
1.1.2(3) [1]

g
g
g

Clauses 1.1(1), 1.1(2)


[3]

Use of Eurocode 8 Part 2 with the other Eurocodes

Part 2 of Eurocode 8 builds on the general provisions of Part 1 (CEN, 2004b) for:
the general performance requirements
seismic action
analysis methods and procedures applicable to all types of structures.

All the general or specic provisions of Part 5 of Eurocode 8 (CEN, 2004a) regarding:
g
g
g
g

siting of the works


properties and seismic verication of the foundation soil
seismic design of the foundation or of earth-retaining structures
seismic soilstructure interaction

apply as well.
Clause 1.2.1 [1,2]
Eurocode 8 is not a stand-alone code. It is applied alongside the other relevant Eurocodes in a
Clauses 1.2.2, 1.2.4 [2] package referring to a specic type of civil engineering structure and construction material.

For bridges, there are four Eurocode packages:


g
g
g
g

2/2:
3/2:
4/2:
5/2:

Concrete bridges
Steel bridges
Composite bridges
Timber bridges.

To be self-sufcient, each package includes all the Eurocode parts needed for design, as
follows:
g

Several EN Eurocodes are included in every single bridge package:


EN 1990: Basis of structural design (including Annex A2: Application for bridges)
EN 1991-1-1: Actions on structures General actions Densities, Self weight and
Imposed loads for buildings
EN 1991-1-3: Actions on structures General actions Snow loads

Chapter 1. Introduction and scope

EN 1991-1-4: Actions on structures General actions Wind actions


EN 1991-1-5: Actions on structures General actions Thermal actions
EN 1991-1-6: Actions on structures General actions Actions during execution
EN 1991-1-7: Actions on structures General actions Accidental actions
EN 1991-2: Actions on structures Trafc loads on bridges
EN 1997-1: Geotechnical Design General rules
EN 1997-2: Geotechnical Design Ground investigation and testing
EN 1998-1: Design of structures for earthquake resistance General rules, seismic
actions, rules for buildings
EN 1998-2: Design of structures for earthquake resistance Bridges
EN 1998-5: Design of structures for earthquake resistance Foundations, retaining
structures, geotechnical aspects.
Additional EN-Eurocodes are included in the Concrete Bridges package (2/2):
EN 1992-1-1: Design of concrete structures General General rules and rules for
buildings
EN 1992-2: Design of concrete structures Concrete bridges Design and detailing
rules.
Additional EN-Eurocodes included in the Steel Bridges package (3/2):
EN 1993-1-1: Design of steel structures General rules and rules for buildings
EN 1993-1-5: Design of steel structures Plated structural elements
EN 1993-1-7: Design of steel structures Strength and stability of planar plated
structures subject to out of plane loading
EN 1993-1-8: Design of steel structures Design of joints
EN 1993-1-9: Design of steel structures Fatigue
EN 1993-1-10: Design of steel structures Selection of steel for fracture toughness and
through-thickness properties
EN 1993-1-11: Design of steel structures Design of structures with tension
components
EN 1993-2: Design of steel structures Steel bridges.
EN-Eurocodes which are included in addition in the Composite Bridges package (4/2) are:
EN 1992-1-1: Design of concrete structures General General rules and rules for
buildings
EN 1992-2: Design of concrete structures Concrete bridges Design and detailing
rules
EN 1993-1-1: Design of steel structures General rules and rules for buildings
EN 1993-1-5: Design of steel structures Plated structural elements
EN 1993-1-7: Design of steel structures Strength and stability of planar plated
structures subject to out of plane loading
EN 1993-1-8: Design of steel structures Design of joints
EN 1993-1-9: Design of steel structures Fatigue
EN 1993-1-10: Design of steel structures Selection of steel for fracture toughness and
through-thickness properties
EN 1993-1-11: Design of steel structures Design of structures with tension
components
EN 1993-2: Design of steel structures Steel bridges
EN 1994-1-1: Design of composite steel and concrete structures General rules and
rules for buildings
EN 1994-2: Design of composite steel and concrete structures General rules and rules
for bridges.

Although package 5/2, for timber bridges, does include Parts 1, 2 and 5 of Eurocode 8, EN 19982:2005 itself is not meant to cover timber bridges.

1.5.

Additional European standards to be used with EN 1998-2:2005

Part 2 of Eurocode 8 makes specic reference to the following product standards:


g
g
g

Clause 1.2.4 [2]

EN 15129:2009: Antiseismic Devices


EN 1337-2:2000: Structural bearings Part 2: Sliding elements
EN 1337-3:2005: Structural bearings Part 3: Elastomeric bearings.
3

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Although EN 1337-5 Structural bearings Part 5: Pot bearings is not specically referenced, it is
also to be used, as relevant.

1.6.
Clauses 1.3(1), 1.3(2)
[1,2]

1.7.
Clause 1.4 [1,2]

Distinction between principles and application rules

Eurocode 8 refers to EN 1990 for the distinction between principles and application rules.
Accordingly, reference is made here also to Designers Guides to other Eurocodes for elaboration. It is noted, though, that, in practice, the distinction between principles and application
rules is immaterial, as all provisions of the normative text are mandatory: non-conformity to a
single application rule disqualies the entire design from being considered to accord with the
EN Eurocodes.

1.8.
Clauses 1.5, 1.6 [2]

Assumptions

Eurocode 8 refers to EN 1990 (CEN, 2002) for general assumptions, so reference is made here to
Designers Guides to other Eurocodes for elaboration. Also, Eurocode 8 adds the condition that
no change to the structure (not even one that increases the force resistance of members) should
take place during execution or afterwards without proper justication and verication.

Terms and denitions symbols

Terms and symbols are dened in the various chapters of this Designers Guide wherever they
rst appear.
REFERENCES

CEN (Comite Europeen de Normalisation) (2002) EN 1990: Eurocode Basis of structural design
(including Annex A2: Application to bridges). CEN, Brussels.
CEN (2004a) EN 1998-5:2004 Eurocode 8 Design of structures for earthquake resistance Part 5:
Foundations, retaining structures, geotechnical aspects. CEN, Brussels.
CEN (2004b) EN 1998-1:2004. Eurocode 8 Design of structures for earthquake resistance Part 1:
General rules, seismic actions and rules for buildings. CEN, Brussels.
CEN (2005a) EN 1998-2:2005 Eurocode 8 Design of structures for earthquake resistance Part 2:
Bridges. CEN, Brussels.
CEN (2005b) EN 1998-3:2005 Eurocode 8 Design of structures for earthquake resistance Part 3:
Assessment and retrotting of buildings. CEN, Brussels.

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.005

Chapter 2

Performance requirements and


compliance criteria
2.1.

Performance-based seismic design of bridges

Paraphrasing for the particular purpose of the seismic design of bridges the b 2010 Model
Code ib, 2012) the seismic performance of a bridge refers to its behaviour under seismic action:
the bridge must be designed, constructed and maintained so that it adequately and in an economically reasonable way performs in earthquakes that may take place during its construction
and service. More specically, the bridge must:
g
g

remain t for the use for which it has been designed


withstand extreme, occasional and frequent seismic actions likely to occur during its
anticipated use and avoid damage by an exceptional earthquake to an extent
disproportionate to the triggering event
contribute positively to the needs of humankind with regard to nature, society, economy
and wellbeing.

Accordingly, three categories of performance are addressed by the b 2010 Model Code
( b, 2012):
g

Serviceability: the ability of the bridge and its structural components to perform, with
appropriate levels of reliability, adequately for normal use after or even during seismic
actions expected during its service life.
Structural safety: the ability of the bridge and its structural components to guarantee the
overall stability, adequate deformability and ultimate load-bearing resistance,
corresponding to occasional, extreme or exceptional seismic actions with appropriate levels
of reliability for the specied reference periods.
Sustainability: the ability of the bridge to contribute positively to the fullment of the
present needs of humankind with respect to nature, society and people, without
compromising the ability of future generations to meet their needs in a similar manner.

In performance-based design, the bridge is designed to perform in a required manner during its
entire life cycle, with performance evaluated by verifying its behaviour against specied requirements, based in turn on stakeholders demands for the bridge performance and required service
life. Performance-based design of a new bridge is completed when it has been shown that the
performance requirements are satised for all relevant aspects of performance related to serviceability, structural safety and sustainability. If the performance of a structure or a structural
component is considered to be inadequate, we say we have failure.
The Eurocodes introduce limit states to carry out performance-based design for serviceability
and safety (CEN, 2002). Limit states mark the boundary between desired and undesirable structural performance of the whole structure or a component: beyond a limit state, one or more performance requirements are no longer met. For the particular case of seismic design, limit states
are dened conceptually for all transient situations in the service life or the execution of the
bridge during which the earthquake acts in combination with any relevant persistent or transient
actions or environmental inuences. They correspond to discrete representations of the structural
response under a specied exposure for which specic losses/damages can be associated. In
practice, they use simplied models for the exposure and the structural response ( b, 2012).
5

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

The Eurocodes recognise (CEN, 2002):


g
g

serviceability limit states (SLSs)


ultimate limit states (ULSs).

SLSs are those beyond which specied requirements for the bridge or its structural components
related to its normal use are no longer met. If they entail permanent local damage or permanent
unacceptable deformations, the outcome of their exceedance is irreversible. It is considered to be
serviceability failure, and may require repair to reinstate tness for use. According to the b 2010
Model Code ( b, 2012), in seismic design at least one but sometimes two SLSs must be
explicitly considered, each one for a different representative value of the seismic action:
g

The operational (OP) limit state: the facility (bridge or any other construction work)
satises the operational limit state criteria if it has suffered practically no damage and can
continue serving its original intention with little disruption of use for repairs; any repair, if
needed, can be deferred to the future without disruption of normal use.
The immediate use (IU) limit state: the facility satises this if all of the following
conditions apply:
the structure itself is very lightly damaged (i.e. localised yielding of reinforcement,
cracking or local spalling of concrete, without residual drifts or other permanent
structural deformations)
the normal use of the facility is temporarily but safely interrupted
risk to life is negligible
the structure retains fully its earlier strength and stiffness and its ability to withstand
loading
the (minor) damage of non-structural components and systems can be easily and
economically repaired at a later stage.

ULSs are limit states associated with the various modes of structural collapse or stages close to it,
which for practical purposes are also considered as a ULS. Exceedance of a ULS is almost always
irreversible; the rst time it occurs it causes inadequate structural safety, that is, failure. ULSs
address (CEN, 2002; b, 2012):
g
g

life safety
protection of the structure.

In seismic design, ULSs that may require consideration include ( b, 2012):


g
g
g
g

reduction of residual resistance below a certain limit


permanent deformations exceeding a certain limit
loss of equilibrium of the structure or part of it, considered as a rigid body (e.g.
overturning)
sliding beyond a certain limit or overturning.

In seismic design there may be several ULSs, with different consequences of limit state failure,
high or medium. According to the b 2010 Model Code ( b, 2012), in seismic design at least
one but normally both of the following ULSs must be explicitly considered, each one for a
different representative value of the seismic action:
g

The life safety (LS) limit state: this is reached if any of the following conditions are met
(but not surpassed):
the structure is signicantly damaged, but does not collapse, not even partly, retaining
its integrity
the structure does not provide sufcient safety for normal use, although it is safe enough
for temporary use
secondary or non-structural components are seriously damaged, but do not obstruct
emergency use or cause life-threatening injuries by falling down
the structure is on the verge of losing capacity, although it retains sufcient load-bearing
capacity and sufcient residual strength and stiffness to protect life for the period until
the repair is completed

Chapter 2. Performance requirements and compliance criteria

repair is economically questionable and demolition may be preferable.


The near-collapse (NC) limit state: this is reached if any of the following conditions are
met:
the structure is heavily damaged and is at the verge of collapse
although life safety is mostly ensured during the loading event, it is not fully guaranteed
as there may be life-threatening injury situations due to falling debris
the structure is unsafe even for emergency use, and would probably not survive
additional loading
the structure presents low residual strength and stiffness but is still able to support the
quasi-permanent loads.

A representative seismic action, with a prescribed probability of not being exceeded during the
design service life, should be dened for each limit state considered. According to the b 2010
Model Code ( b, 2012), multiple representative seismic actions appropriate for ordinary facilities
are:
g

g
g
g

For the operational (OP) limit state: a frequent seismic action, expected to be exceeded at
least once during the design service life (i.e. having a mean return period much shorter
than the design service life).
For immediate use (IU): an occasional earthquake, not expected to be exceeded during
the design service life (e.g. with a mean return period about twice the design service life).
For life safety (LS): a rare seismic action, with a low probability of being exceeded (10%)
during the design service life.
For near-collapse (NC): a very rare seismic action, with very low probability of being
exceeded (25%) in the design service life of the structure.

For facilities whose consequences of failure are very high, the very rare seismic action may be
appropriate for the life safety limit state. For those which are essential for the immediate postearthquake period, a rare seismic action may be appropriate for the immediate use or even
the operational limit state ( b, 2012).
A fully edged performance-based seismic design of a bridge as outlined above for the case of the
b 2010 Model Code ( b, 2012) will serve well the interests and objectives of owners, in that it
allows explicit verication of performance levels related to different level of operation (including
loss) of the bridge under frequent, occasional, rare or quite exceptional earthquakes. However,
the design process may become too complex and cumbersome. Therefore, even the b 2010
Model Code ( b, 2012) recognises that, depending on the use and importance of the facility,
competent authorities will choose how many and which limit states should be veried at a
minimum and which representative seismic action they will be paired with. The seismic design
of a bridge, or at least certain of its aspects, may be conditioned by just one of these limit
states. However, this may hold on a site-specic but not on a general basis, because the seismicity
of the site controls the relative magnitude of the representative seismic actions for which the
multiple limit states should be veried.
In closing this discussion on the performance-based design of bridges, a comment is required on
sustainability performance: it is not explicitly addressed in the rst generation of Eurocodes, but
will be in the next one, as the European Union recently added Sustainable use of resources
to the two essential requirements of Mechanical resistance and stability and Resistance to
re for construction products that must be served by the Eurocodes. The b 2010 Model
Code ( b, 2012), which has raised sustainability performance to the same level as serviceability
and structural safety, speaks about it still in rather general terms. At any rate, the sustainabilityconscious bridge designer should cater in the conceptual design phase for aesthetics and the
minimisation of environmental impact (including during execution) and during all phases,
from concept to detailed design, for savings in materials.

2.2.

Performance requirements for new bridges in Eurocode 8

Part 2 of Eurocode 8 (CEN, 2005) requires a single-level seismic design of new bridges with the
following explicit performance objective:

Clauses 2.1(1),
2.2.2(1), 2.2.2(4) [2]
7

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

The bridge must retain its structural integrity and have sufcient residual resistance to be
used for emergency trafc without any repair after a rare seismic event the design
seismic action explicitly dened in Parts 1 and 2 of Eurocode 8; any damage due to this
event must be easily repairable.

Although called a non-collapse requirement, in reality this corresponds to the life safety, rather
than to the near-collapse, limit state of the general framework of performance-based seismic
design outlined in the previous section, since sufcient residual resistance has to be available
after the design seismic event for immediate use by emergency trafc.
Clause 2.1(1) [1,2]

As we will see in more detail in Section 3.12.2 of this Guide, the design seismic action of
structures of ordinary importance is called the reference seismic action; its mean return
period is the reference return period, denoted by TNCR . Eurocode 8 recommends basing
the determination of the design seismic action on a 10% exceedance probability in 50 years,
corresponding to a reference return period of 475 years.

Clause 2.2.2(5) [2]

If the seismicity is low, the probability of exceedance of the design seismic action during the
design life of the bridge may be well below 10%, and at any rate difcult to quantify. For
such cases, Eurocode 8 allows for consideration of the seismic action as an accidental action;
also, in these cases it tolerates more damage to the bridge deck and secondary components, as
well to the bridge parts intended for controlled damage under the design seismic action.

Clauses 2.1(2)2.1(6)
[2]
Clause 3.2.1(3) [1]

Again as detailed in Section 3.12.2 of this Guide, Eurocode 8 pursues enhanced performance for
bridges that are vital for communications in the region or very important for public safety, not by
upgrading the performance level, as suits the general framework of performance-based seismic
design delineated in the previous section, but by modifying the hazard level (increasing the
mean return period) for the design seismic action under which the non-collapse requirement
is met. This is done by multiplying the reference seismic action by the importance factor gI ,
which by denition is gI 1.0 for bridges of ordinary importance (i.e. for the reference return
period of the seismic action).

Clauses 2.2.1(1),
2.2.3(1), 2.3.1(1) [2]

Part 2 of Eurocode 8 calls also for the limitation of damage under a loosely dened seismic action
with a high probability of exceedance; such damage must be minor and limited only to secondary
components and to the parts of the bridge intended for controlled damage under the design
seismic action. However, this requirement is of no practical consequence for design: it is
presumed to be implicitly fullled if all the criteria for compliance with the non-collapse
requirement above are checked and met. This should be contrasted with new buildings, for
which Part 1 of Eurocode 8 (CEN, 2004) provides explicit checks under a well-dened
damage limitation seismic action. However, these damage checks (inter-storey drifts)
normally refer to non-structural elements that are not present in bridges.

Clauses 2.3.4(1),
2.3.4(2) [2]

Although not explicitly stated, an additional performance requirement for bridges designed to
face the design seismic action by means of ductility and energy dissipation is the prevention
of the near-collapse limit state in an extreme and very rare, as yet undened, earthquake.
This implicit performance objective is pursued through systematic and across-the-board application of the capacity design concept, which allows full control of the inelastic response
mechanism.

2.3.
Clauses 2.2.2(3),
2.3.2.2(4) [2]

Clause 2.2.2(5) [2]

Compliance criteria for the non-collapse requirement and


implementation

2.3.1
Design options to meet the bridge performance requirements
For continued use after the design seismic action (e.g. by emergency trafc), the deck of the
bridge must remain in the elastic range. Damage should be local and limited to non-structural
or secondary components, such as expansion joints, parapets or concrete slabs providing topslab continuity between adjacent simply-supported spans, most often built of precast concrete
girders. The latter may yield during bending of the deck in the transverse direction.
If the seismic action is considered in the National Annex as accidental, because the probability
of exceedance of the design seismic action during the design life of the bridge is well below 10%

Chapter 2. Performance requirements and compliance criteria

or undened, Eurocode 8 allows as an exemption some inelastic action in and damage to the
bridge deck.
It is today commonplace that the earthquake represents for the structure a demand to accom- Clauses 2.4(3), 2.4(4),
6.6.2.3(1) [2]
modate imposed dynamic displacements primarily in the horizontal direction and not
forces. Seismic damage results from them. The prime aim of seismic design is to accommodate
these horizontal displacements with controlled damage. The simple structural system of
bridges lends itself to the following options:
To place the deck on a system of sliding or horizontally exible bearings (or bearing-type
devices) at the top of the substructure (the abutments and all piers) and accommodate the
horizontal displacements at this interface.
2 To x or rigidly connect horizontally the deck to the top of at least one pier but let it slide
or move on exible bearings at all other supports (including the abutments). The piers that
are rigidly connected to the deck are required to accommodate the seismic horizontal
displacements by bending. These piers develop inelastic rotations in exural plastic
hinges, if they are not tall and exible enough to accommodate the horizontal
displacements elastically.
3 To accommodate (most of ) the seismic horizontal displacements in the foundation and the
soil, either through sliding at the base of piers or through inelastic deformations of soil
pile systems of the foundation.
4 To rigidly connect the deck with the abutments (either monolithically or via xed bearings
or links) into an integral system that follows the ground motion with little additional
deformation of its own. It then makes little difference if any intermediate piers are also
integral with the deck or support it on bearings.
1

Option 4 (usually termed integral bridges) is encountered only in relatively short bridges with
one or very few spans. It is dealt with in Section 5.4 of this Designers Guide as a special case.

Clauses 4.1.6(9),
4.1.6(10) [2]

Part 5 of Eurocode 8 explicitly allows horizontal sliding of footings with respect to the soil (as
long as residual rotation about horizontal axes and overturning are controlled), but this is an
unconventional design option adopted for major bridges, notably the 2.45 km continuousdeck Rion-Antirrion bridge with a design ground acceleration of 0.48g. For typical bridges, a
non-reversible sliding of one foundation support may entail serious problems. Part 5 of
Eurocode 8, as well as Part 2, also allows inelastic deformations in foundation piles. This may
be the only viable option if the deck is monolithic with strong and rigid wall-like piers placed
transverse to the bridge axis.

Clauses 5.4.1.1(7),
5.4.2(7) [3]
Clause 4.1.6(7) [2]

Most common in practice are options 1 and 2, which are therefore considered as the two fundamental options for the seismic design of bridges. Option 1 is considered in Part 2 of Eurocode 8
as full seismic isolation, with the piers designed to remain elastic during the design seismic
action.

Clauses 2.3.2.1(10),
4.1.6(11) [2]

In option 2, the piers are normally designed to respond well into the inelastic range, mobilising
ductility and energy dissipation to withstand the seismic action. Design based on ductility and
energy dissipation capacity is seismic design par excellence. It is at the core of Part 2 of
Eurocode 8, where it is called design for ductile behaviour, as well as of this Designers Guide.

Clauses 2.2.2(2),
2.2.2(4), 2.3.2.2(1),
2.3.2.2(2), 2.3.2.2(7),
4.1.6(6) [2]

Ductility and energy dissipation under the design seismic action is entrusted by Part 2 of
Eurocode 8 to the piers, and is understood to entail a certain degree of damage at the plastic
hinges (spalling of the unconned concrete shell outside the conning hoops, but no buckling
or fracture of bars, nor crushing of conned concrete inside the hoops). However, as this
damage is meant to be reparable, it should be limited to easily accessible parts of the pier.
Parts above the normal water level (be it in a sleeve or casing) are ideal. Those at a shallow
depth below grade but above the normal water table are also accessible. Those embedded
deeper in ll but above the normal water level are still accessible but with increased difculty.
Part 2 of Eurocode 8 does not distinguish in great detail between these cases. It considers,
though, as accessible the base of a pier deep in backll but as inaccessible parts of the pier
which are deep in water, or piles under large pile-caps; to reduce damage in such regions

Clauses 2.2.2(4),
2.3.2.2(3) [2]

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

under the design seismic action, it divides seismic design forces by 0.6 should plastic hinges form
there.
2.3.2
Design of bridges for energy dissipation and ductility
2.3.2.1 Introduction
Section 2.3.2 refers to one of the two fundamental options for the seismic design of bridges,
namely to option 2: that of xing horizontally the deck to the top of at least one pier but to
let it slide at the abutments and accommodate the seismic horizontal displacements through
bending of the piers, with ductile and dissipative exural plastic hinges forming at their
ends.

Clauses 2.3.5.2(1),
2.3.5.2(2), 2.3.6.1(8)
[2]

2.3.2.2 Design of the bridge as a whole for energy dissipation and ductility
It has already been pointed out that the earthquake is a dynamic action, representing for a structure a requirement to sustain certain displacements and deformations and not specic forces.
Eurocode 8 allows bridges to develop signicant inelastic deformations under the design
seismic action, provided that the integrity of individual components and of the bridge as a
whole is not jeopardised. Design of a bridge to Eurocode 8 for the non-collapse requirement
under the design seismic action is force-based, nonetheless.
The foundation of force-based seismic design for ductility and energy dissipation is the inelastic
response spectrum of a single-degree-of-freedom (SDoF) system having an elasticperfectly
plastic forcedisplacement curve, F  d, in monotonic loading. For given period, T, of the
elastic SDoF system, the inelastic spectrum relates:
g
g

the ratio q Fel/Fy of the peak force, Fel , that would have developed if the SDoF system
were linear elastic, to the yield force of the system, Fy
the maximum displacement demand of the inelastic SDoF system, dmax , expressed as ratio
to the yield displacement, dy (i.e. as the displacement ductility factor, md dmax/dy).

Part 2 of Eurocode 8 has adopted a modication of the inelastic spectra proposed in Vidic et al.
(1994):

md q
md 1 q  1
md 1

1:25TC
 5q  4
T

if T  1:25TC

D2:1a

if T , 1:25TC

D2:1b

if T , 0:033 s

D2:1c

where TC is the transition period of the elastic spectrum, between its constant spectral pseudoacceleration and constant spectral pseudo-velocity ranges (see Section 3.1.3). Equation (D2.1)
expresses Newmarks well-known equal displacement rule; that is, the empirical observation
that in the constant spectral pseudo-velocity range the peak displacement response of the
inelastic and of the elastic SDoF systems are about the same.
With F being the total lateral force on the structure (the base shear, if the seismic action is in the
horizontal direction), the ratio q Fel/Fy is termed in Eurocode 8 the behaviour factor (the
force reduction factor or the response modication factor, R, in North America). It is
used as a universal reduction factor on the internal forces that would develop in the elastic
structure for 5% damping, or, equivalently, on the seismic inertia forces that would develop
in this elastic structure and cause, in turn, the seismic internal forces. In this way, the seismic
internal forces for which the members of the structure should be dimensioned can be calculated
through linear-elastic analysis. In return, the structure must be provided with the capacity to
sustain a peak global displacement at least equal to its global yield displacement multiplied
by the displacement ductility factor, md , that corresponds to the value of q used for the
reduction of elastic force demands (e.g. according to Eqs (D2.1)). This is termed the ductility
capacity, or the energy-dissipation capacity as it has to develop through cyclic response in
which the members and the structure as a whole dissipate part of the seismic energy input
through hysteresis.
10

Chapter 2. Performance requirements and compliance criteria

2.3.2.3 Design of plastic hinges for energy dissipation and ductility


In force-based seismic design for ductility and energy dissipation, exural plastic hinges in piers
are dimensioned and detailed to achieve a combination of force resistance and ductility that
provides a safety factor between 1.5 and 2 against substantial loss of resistance to lateral (i.e.
horizontal) load. To this end, they are rst dimensioned to provide a design value of moment
and axial force resistance Rd , at least equal to the corresponding action effects due to the
seismic design situation, Ed , from the analysis:
Ed  Rd

Clause 2.3.3(1) [2]

(D2.2)

The values of Ed in Eq. (D2.2) are due to the combination of the seismic action with the
quasi-permanent gravity actions (i.e. the nominal permanent loads and the quasi-permanent
trafc loads, as pointed out in Section 5.4 in connection with Eq. (D5.6a) for the calculation
of the deck mass). As linear analysis is normally applied, Ed may be found from superposition
of the seismic action effects from the analysis for the seismic action alone to the action effects
from that for the quasi-permanent gravity actions.
After having been dimensioned to meet Eq. (D2.2), exural plastic hinges in piers are detailed to
provide the deformation and ductility capacity necessary to meet the deformation demands on
them from the design of the structure for the chosen q-factor value. The measure used for the
deformation and ductility capacity of exural plastic hinges is the curvature ductility factor of
the pier end section, whose supply-value is

mf fu/fy

Clauses 2.3.5.1(1),
2.3.5.3(1), 2.3.5.3(2),
2.3.6.1(8), Annex B,
Annex E [2]

(D2.3)

where fy is the yield curvature of that section (computed from rst principles) and fu its ultimate
curvature (again from rst principles and the ultimate deformation criteria adopted for
the materials). At the other end, the global displacement demands are expressed through the
global displacement ductility factor of the bridge, md , connected to the q factor used in the
design of the bridge through the inelastic spectra, in this case Eqs (D2.1).
The intermediary between mf and md is the ductility factor of the chord rotation at the pier end
where the plastic hinge forms, mu . Recall that the chord rotation u at a pier end is the deection
of the inexion point with respect to the tangent to the pier axis at the end of interest, divided by
the distance between these two points of the pier, termed the shear span and denoted by Ls . So,
the chord rotation u is a measure of member displacement, not of the relative rotation between
sections. If the pier is xed at its base against rotation and supports the deck without the intervention of horizontally exible bearings (i.e. if it is monolithically connected or supported on
the pier through xed e.g. pot bearings), the chord rotation at the hinging end of the pier
is related as follows to the deck displacement right above the pier top, d, in the common
cases of:
Pier columns monolithically connected at the top to a very stiff deck with near-xity there
against rotation for seismic response in the longitudinal direction and inexion point at the
column mid-height (see Section 5.4, Eqs (D5.4) if the deck cannot be considered as rigid
compared with the piers in the longitudinal direction); the horizontal displacement of that
point is one-half of that of the deck above, d, and the shear span, Ls , is about equal to
one-half of the pier clear height, Hp ; Ls  Hp/2; therefore, at the plastic hinges forming at
both ends of the pier, u  0.5d/Ls d/Hp .
2 Multiple-column piers monolithically connected at the top to a very stiff deck or a cap
beam with near-xity there against rotation for seismic response in the transverse
direction; the situation is similar to case 1 above, so in the transverse direction Ls  Hp/2
and u  0.5d/Ls d/Hp .
3 Piers supporting the deck through xed (e.g. pot) bearings at the top and working as
vertical cantilevers with a shear span Ls about equal to the pier clear height, Ls  Hp and
u d/Ls d/Hp .
4 Single-column piers monolithic with the deck and working in the transverse direction of
the bridge as vertical cantilevers; if the rotational inertia of the deck about its longitudinal
axis and the vertical distance between the pier top and the point of application of the deck
1

11

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

inertia force are neglected (see Section 5.4, Eqs (D5.5) for the case they are not), the
situation is similar to case 3 above, so in the transverse direction Ls  Hp and
u d/Ls d/Hp .
In a well-designed bridge, all piers will yield at the same time, turning the bridge into a fully
edged plastic mechanism. Then, in all cases 1 to 4 above, md d/dy will be (about) equal to
mu u/uy:

md  mu

(D2.4)

In a plastic hinge model of the inelastic deformation of the pier, all inelastic deformations are
lumped in the plastic hinge, which is considered to have a nite length Lpl and to develop
constant inelastic curvature all along Lpl . Then, for a linear bending moment diagram
(constant shear force) along the pier, the chord rotation at its yielding end(s) is



Lpl
Ls 
u fy f  fy Lpl 1 
3
2Ls

D2:5

giving, for uy fyLs/3 (purely exural elastic behaviour),




 3Lpl
Lpl
mu 1 mf  1
1
Ls
2Ls


D2:6

Equation (D2.6) is inverted as

mf

fu
md  1
1
fy
3l1  0:5l

D2:7

where l Lpl/Ls . Then, if the pier plastic hinge length Lpl is estimated through appropriate
empirical relations, Eqs (D2.1) and (D2.7) translate the q factor used in the design of the
bridge into a demand value for the curvature ductility factor of the piers. Note that Part 1 of
Eurocode 8 has adopted for concrete members in buildings the following conservative approximation of Eqs (D2.6) and (D2.7):

mu 1 0.5(mf  1)

i.e. mf 2mu  1

(D2.8)

dating from the ENV version of the concrete buildings part of Eurocode 8 (ENV 1998-1-3:1994).
Clauses 2.3.5.4(1),
2.3.5.4(2) [2]

Clauses 2.3.3(1),
2.3.4(1), 2.3.4(2),
2.3.6.2(2) [2]

If linear analysis is used alongside the design spectrum involving the q factor, the required value
of the curvature ductility factor of the piers is presumed to be provided if the detailing rules of
Part 2 of Eurocode 8 are applied, prescriptive or not. If nonlinear analysis is used instead, the
inelastic chord rotation demands obtained from it are compared with appropriate design
values of chord rotation capacities, obtained by setting f fu in Eq. (D2.5). Details are given
in Chapter 6.
2.3.2.4 Capacity design for the ductile global response
The bridges seismic design determines how the (roughly) given peak global displacement
demand of the design seismic action is distributed to its various components. Eurocode 8 uses
capacity design to direct and limit this demand only to those best suited to withstand it.
Capacity design imposes a hierarchy of strengths between adjacent components or regions, and
between different mechanisms of load transfer in the same member, so that those items capable of
ductile behaviour and hysteretic energy dissipation are the rst ones to develop inelastic deformations. More importantly, they do so in a way that precludes the development of inelastic
deformations in any component, region or mechanism deemed incapable of ductile behaviour
and hysteretic energy dissipation.
The components, regions thereof or mechanisms of force transfer to which the peak global
displacement and deformation energy demands are channelled by capacity design are selected,

12

Chapter 2. Performance requirements and compliance criteria

taking into account the following aspects:


Their inherent ductility. Ductile components, regions thereof or mechanisms of force
transfer are entrusted through capacity design for inelastic deformations and energy
dissipation, while brittle ones are shielded from them. Flexure is a far more ductile
mechanism of force transfer than shear, and can be made even more so through judicious
choice of the level of axial force and the amount, distribution and ductility of longitudinal
and transverse reinforcement.
2 The role of the component for the integrity of the whole and the fullment of the
performance requirements of the bridge. The foundation and connections between
components (bearings, links, holding-down devices, etc.) securing structural integrity are
most important for the stability and integrity of the whole; the integrity of the deck itself
determines the continued operation of the bridge after the earthquake.
3 Accessibility and difculty in inspecting and repairing any damage. Accessible regions of
the piers (above the grade and the water level) are the easiest to repair without disruption
of trafc.
1

On the basis of the above aspects, a clear hierarchy of the bridge components and mechanisms of
force transfer emerges, determining the order in which they are allowed to enter the inelastic
range during the seismic response: the deck, the connections between components and the
foundation are to be shielded from inelastic action; the last is channelled to exural plastic
hinges at accessible ends of the piers. Capacity design ensures that this order is indeed respected.
As we will see in more detail later, it works as follows.
The required force resistance of the components, regions thereof or mechanisms of force
transfer to be shielded from inelastic response is not determined from the analysis. Simple
calculations (normally on the basis of equilibrium alone) are used instead, assuming that
all relevant plastic hinges develop their moment resistances in a way that prevents preemptive attainment of the force resistance of the components, etc., to be shielded from
inelastic action.
2.3.2.5 How elastic deformations in exible bearings or the foundation ground
affect the ductility of the bridge
Assume that the deck is supported on a ductile pier that can develop a curvature ductility factor
mfo in the plastic hinge(s) and a chord rotation ductility factor muo , and has elastic lateral
stiffness Kp if xed at the base:

Annex B [2]

(a) For a single-column pier presenting exural rigidity (EI )c in a vertical plane in the
transverse direction of the bridge and supporting a deck mass with a radius of gyration
rm,d about its centroidal axis (r2m,d ratio of tributary rotational mass moment of inertia
of the deck about the decks centroidal axis to the tributary deck mass):
"

Kp 
H

9r2m;d
8Ls

3EIc
!
#


2
Ls H ycg ycg

D2:9a

In Eq. (D2.9a), ycg is the distance from the soft of the deck to the centroid of its
section, and Ls is the shear span at the pier base (see Eq. (D5.5) in Section 5.4 for this
particular case).
(b) For a pier consisting of n  1 columns, each one with height H and presenting the rigidity
(EI )c within the plane of bending considered, all having the top xed to the soft of a
very stiff deck:
Kp 12Sn(EI )c/H3

(D2.9b)

Single-column piers (n 1) in the transverse direction of the bridge are not addressed by
this case but by case 1 above and Eq. (D2.9a).
13

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

(c) For a pier as in case 2 above but with the top of its n  1 columns pin-connected to the
deck, instead of being xed to it:
Kp 3Sn(EI )c/H3

(D2.9c)

Assume also that one or more additional components intervene between the deck and the ground
in series with the pier, all designed to remain elastic until and after the pier yields (e.g. through
capacity design according to the previous section). The generic elastic stiffness of these
components is denoted as Kel . Such components can be:
g

the compliance of the ground if the foundation itself has horizontal stiffness Kfh (base
shear divided by the horizontal displacement of the foundation) and rotational stiffness
Kff (moment divided by the rotation at the pier base), giving horizontal stiffness at the top
of the pier Kff/H2 and/or
an elastic (e.g. elastomeric) bearing with horizontal stiffness Kb (Kb GA/t if the bearing
has horizontal section area A, and its material the elastomer has total thickness t and
shear modulus G) needless to say, this case does not combine with piers of case 2 above,
which have their top xed to the soft of the deck.

The deck sees down below a total stiffness K such that


X 1
1
1
1
1
1
H2

K Kp
Kel Kp Kb Kfh Kff

D2:10

If the pier yields at a base shear Vy , the displacement of the deck at yielding is

dy

Vy Vy X Vy

K
Kp
Kel

D2:11

After the pier yields, additional horizontal displacements are due to the inelastic rotation(s) of its
plastic hinge(s) alone, giving an inelastic displacement of the deck:

Vy
md dy muo  1
dy
Kp

D2:12

where md is the global displacement ductility factor of the bridge at the level of the deck.
According to Eqs (D2.4), (D2.6) and (D2.7), there is proportionality between (md  1),
(mu  1) and (mf  1). Therefore, Eqs (D2.9)(D2.11) state that, to achieve the same target
value md of the global displacement ductility factor of the bridge at the level of the deck, the
curvature ductility demand at a plastic hinge of a pier should increase from mfo to

X Kp 


mf 1 mfo  1 1
Kel

D2:13

The horizontal stiffness of an elastic bearing is several times smaller than that of an ordinary pier.
So, Eq. (D2.13) gives unduly large values of the curvature ductility demand that a plastic hinge in
the pier may have to bear for the bridge to achieve the q factor values normally used in the
seismic design of bridges for ductility. So, if piers are intended to resist the design seismic
action through ductility and energy dissipation, elastic bearings have no place on top of them.
By the same token, ductile piers should be nearly xed to the ground: compliance of the foundation will penalise detailing of their plastic hinges for the target q factor of seismic design for
ductility.
The same conclusion can be reached through energy considerations. The pier is an assembly of
components in series, only one of which (the pier shaft having stiffness Kp) possesses the
capability of hysteretic energy dissipation. The other components (elastic bearings at the pier
top and foundation compliance, having a composite stiffness Kel) should remain within the
elastic range. As the same shear force acts on all components in series, the strain energy
14

Chapter 2. Performance requirements and compliance criteria

input in each component at any instant of the seismic response is proportional to their
exibilities, 1/Kp and 1/Kel , respectively. When the portion of the energy input in the dissipative
component is small compared with the input in the series system, the dissipation capability of
the whole assembly is also small. In other words, the behaviour of such assemblies becomes
practically elastic.
2.3.3

Seismic design of bridges for strength instead of ductility: limited ductile


behaviour
Part 2 of Eurocode 8 gives the option to design a bridge to resist the seismic action through
strength alone, without explicitly resorting to ductility and energy dissipation capacity. In this
option, the bridge is designed:
g
g
g

Clauses 2.3.2.1(1),
2.3.2.2(1), 2.3.2.3(1),
2.3.3(2), 2.3.4(3),
2.3.5.4(3) [2]

in accordance with Eurocodes 2, 3, 4 and 7, with the seismic action considered as a static
loading (like wind)
without capacity design considerations, except for non-ductile connections or structural
components (xed bearings, sockets and anchorages of cables and stays), but
observing:
some minimum requirements for the ductility of steel reinforcement or steel sections and
for connement and bar anti-buckling restraint in potential plastic hinges of concrete piers
simplied rules for the ULS verication in shear.

The seismic lateral forces are derived from the design response spectrum using a behaviour
factor, q, not higher than the value of 1.5 attributed to material overstrength. In fact:
(a) if the bridge seismic response is dominated by upper modes (as in cable-stayed bridges) or
(b) concrete piers have:
axial force ratio hk Nd/Ac fck (axial load due to the design seismic action and the
concurrent gravity loads, Nd , normalised to product of the pier section area and the
characteristic concrete strength, Ac fck), higher than or equal to 0.6, or
shear-span ratio, Ls/h, in the direction of bending less than or equal to 1.0,

Clauses 2.3.2.3(2),
4.1.6(3), 4.1.6(5) [2]

then the behaviour factor, q, is taken equal to 1.0.


As design seismic forces are derived with a value of the behaviour factor, q, possibly greater than
1.0, structures designed for strength and little engineered ductility and energy dissipation
capacity are termed limited ductile, in lieu of non-ductile.
Part 2 of Eurocode 8 recommends (in a note) designing the bridge for limited-ductile behaviour
in cases of low seismicity (see below), but does not discourage the designer from using this
option in other cases as well. It species the option as the only possible one, no matter
whether the bridge is a low-seismicity case or not, in two very specic but also quite common
cases:

Clauses 2.3.7(1),
2.3.2.1(1), 2.4(2),
2.4(3), 4.1.6(3),
4.1.6(9)4.1.6(11) [2]

when the deck is fully supported on a system of sliding or horizontally exible bearings (or
bearing-type devices) at the top of the substructure (the abutments and all piers), which
accommodate the horizontal displacements (see option 1 in Section 2.3.1 and the inuence
of non-dissipative components in Section 2.3.2.5 above), or
2 when the deck is rigidly connected to the abutments, monolithically or via xed bearings
or links (listed as option 4 in Section 2.3.1 of this Guide).
1

2.3.4
The balance between strength and ductility
The option described in Section 2.3.3 above, namely to design for strength alone without
engineered ductility and energy dissipation capacity, is an extreme, specied by Part 2 of
Eurocode 8 only for cases a and b and 1 and 2 well delineated in Section 2.3.3. Outside of
these specic cases, the designer is normally given the option to opt for more strength and less
ductility (i.e. for limited-ductile behaviour) or vice versa (for ductile behaviour).

Clause 2.3.2.1(1)

Equations (D2.1) show that, except for short-period bridges, the magnitude of the design seismic
forces decreases when the global displacement ductility factor, md , increases. So, there is an
15

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

apparent economic incentive to increase the available global ductility, to reduce the internal
forces for which the components of the bridge are dimensioned. Moreover, if the lateral force
resistance of the bridge is reduced, by dividing the elastic lateral force demands by a high qfactor value, the verication of the foundation soil, which is done for strength rather than for
ductility and deformation capacity, is much easier. Last but not least, a bridge with ample
ductility supply is less sensitive to the magnitude and the details of the seismic action, and, in
view of the large uncertainty associated with the extreme seismic action in its lifetime, may be
a better earthquake-resistant design.
On the other hand, there are strong arguments for less ductility and dissipation capacity in
seismic design and more lateral force resistance instead. Ductility necessarily entails damage.
So, the higher the lateral strength of the bridge, the smaller will be the structural damage, not
only during more frequent, moderate earthquakes but also due to the design seismic action.
From the construction point of view, detailing piers for more strength is much easier and
simpler than detailing for higher ductility. Also, some bridge congurations may impart signicant lateral force resistance. In others (notably when the deck is rigidly supported on tall and
exible piers), the dominant vibration modes may fall at the long-period tail of the spectrum,
where design spectral accelerations may be small even for q  1.5, and dimensioning the piers
for the resulting lateral force resistance may be trivial. Last but not least, if the bridge falls
outside the framework of common structural congurations mainly addressed by Eurocode 8
(e.g. as in arch bridges, or those having some inclined piers or piers of very different height,
especially if the height does not increase monotonically from the abutments to mid-span), the
designers may feel more condent if they narrow the gap between the results of the linearelastic analysis, for which members are dimensioned and the nonlinear seismic response under
the design seismic action (i.e. if q  1.5 is used).
Clauses 3.2.1(4) [1]
Clause 2.3.7(1) [2]

2.3.5
The cases of low seismicity
Eurocode 8 recommends in a note designing the bridge for limited ductile behaviour if it falls in
the case of low seismicity. Although it leaves it to the National Annex to decide which combination of categories of structures, ground types and seismic zones in a country correspond
to the characterisation as cases of low seismicity, it recommends in a note as a criterion
either the value of the design ground acceleration on type A ground (i.e. on rock), ag (which
includes the importance factor gI), or the corresponding value, agS, over the ground type of
the site (see Section 3.1.2.3 of this Guide for the soil factor, S). Moreover, it recommends a
value of 0.08g for ag , or of 0.10g for agS, as the threshold for the low-seismicity cases.

Clauses 2.2.1(4),
3.2.1(5) [1]

2.4.

Exemption from the application of Eurocode 8

Eurocode 8 itself states that its provisions need not be applied in cases of very low seismicity.
As in cases of low seismicity, it leaves it to the National Annex to decide which combination
of category of structures, ground types and seismic zones in a country qualify as cases of very
low seismicity. It does recommend in a note, though, the same criterion as for the cases of
low seismicity: either the value of the design ground acceleration on type A ground (i.e. on
rock), ag , or the corresponding value, agS, over the ground type of the site. It recommends a
value of 0.04g for ag , or of 0.05g for agS, as the threshold for the very low seismicity cases.
Because the value of ag includes the importance factor gI , ordinary bridges in a region may
be exempted from the application of Eurocode 8, while more important ones may not be.
This is consistent with the notion that the exemption from the application of Eurocode 8 is
due to the inherent lateral force resistance of any structure designed for non-seismic loadings,
neglecting any contribution from ductility and energy dissipation capacity. Given that
Eurocode 8 considers that, because of overstrength, any structure is permitted a behaviour
factor, q, of at least 1.5, implicit in the value of 0.05g for agS recommended as the ceiling for
very low seismicity cases is a presumed lateral force capacity of 0.05  2.5/1.5 0.083g.
If a National Annex states that the entire national territory is considered as a case of very low
seismicity, then Eurocode 8 (all six parts) does not apply at all in the country.
REFERENCES

CEN (Comite Europeen de Normalisation) (2002) EN 1990: Eurocode Basis of structural design
(including Annex A2: Application to bridges). CEN, Brussels.
16

Chapter 2. Performance requirements and compliance criteria

CEN (2004) EN 1998-1:2004: Eurocode 8 Design of structures for earthquake resistance Part 1:
General rules, seismic actions and rules for buildings. CEN, Brussels.
CEN (2005) EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance Part 2:
Bridges. CEN, Brussels.
b (2012) Model Code 2010, vol. 1. b Bulletin 65. Federation Internationale du Beton, Lausanne.
Vidic T, Fajfar P and Fischinger M (1994) Consistent inelastic design spectra: strength and
displacement. Earthquake Engineering and Structural Dynamics 23: 502521.

17

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.019

Chapter 3

Seismic actions and geotechnical


aspects
3.1.

Design seismic actions

3.1.1
Introduction
As pointed out in Section 2.2 of this Guide, Eurocode 8 entails a single-tier seismic design of new
bridges, with verication of the no-(local-)collapse requirement under the design seismic action
alone. So, whatever is said in Section 3.1 refers to the design seismic action of the bridge. Note,
however, that in its two-tier seismic design of buildings and other structures, Part 1 of Eurocode 8
(CEN, 2004a) adopts the same spectral shape for the different seismic actions to be used for
different performance levels or limit states. The difference in the hazard level is reected only
through the peak ground acceleration to which each spectrum is anchored.
The seismic action is considered to impart concurrent translational motion in three orthogonal
directions: the vertical and two horizontal ones at right angles to each other. The motion is
taken to be applied at the interface between the structure and the ground. If springs are used
to model the soil compliance under and/or around spread footings, piles or shafts (caissons),
the motion is considered to be applied at the soil end of these springs.
3.1.2
Elastic response spectra
3.1.2.1 Introduction
The reference representation of the seismic action in Eurocode 8 is through the response
spectrum of an elastic single-degree-of-freedom (SDoF) oscillator having a given viscous
damping ratio (with 5% being the reference value). Any other alternative representation of the
seismic action (e.g. in the form of acceleration time histories) should conform to the elastic
response spectrum for the specied value of the damping ratio.

Clauses 3.1.1(2),
3.1.2(1)3.1.2(3) [2]

Clause 3.2.2.1(1) [1]


Clause 3.2.1(1) [2]

Clause 7.5.4(3),
Table 7.1 [2]

Because:
g

Clause 2.1(1) [2]


Clauses 2.1(1),
2.2.1(1) [1]

earthquake ground motions are traditionally recorded as acceleration time histories and
seismic design is still based on forces, conveniently derived from accelerations,

the pseudo-acceleration response spectrum, Sa(T ), is normally used. If spectral displacements,


Sd(T ), are of interest, they can be obtained from Sa(T ), assuming simple harmonic oscillation:
Sd(T ) (T/2p)2Sa(T )

(D3.1)

Spectral pseudo-velocities can also be obtained from Sa(T ) as


Sv(T ) (T/2p)Sa(T )

(D3.2)

Note that pseudo-values do not correspond to the real peak spectral velocity or acceleration. For
a damping ratio of up to 10% and for a natural period T between 0.2 and 1.0 s, the pseudovelocity spectrum closely approximates the actual relative velocity spectrum.
3.1.2.2 Design ground accelerations importance classes for reliability differentiation
Clauses 3.2.1(2),
In Eurocode 8 the elastic response spectrum is taken as proportional (anchored) to the peak
3.2.2.1(1),3.2.2.3(1)[1]
acceleration of the ground:
19

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

g
g

Clause 2.1(3) [2]


Clause 3.2.1(3) [1]

the horizontal peak acceleration, ag , for the horizontal component(s) of the seismic action
or
the vertical peak acceleration, avg , for the vertical component.

The basis of the seismic design of new bridges in Eurocode 8 is the design seismic action, for
which the no-(local-)collapse requirement should be met. It is specied through the design
ground acceleration in the horizontal direction, ag , which is equal to the reference peak
ground acceleration on rock from national zonation maps, multiplied by the importance
factor, gI , of the bridge:
ag gIagR

(D3.3)

For bridges of ordinary importance (belonging to importance class II in Eurocode 8), by


denition gI 1.0.
Clauses 2.1(4)2.1(6)
[2]

Eurocode 8 recommends classifying in importance class III any bridge that is crucial for communications, especially in the immediate post-earthquake period (including access to emergency
facilities), or whose downtime may have a major economic or social impact, or which by failing
may cause large loss of life, as well as major bridges with a target design life longer than the
ordinary nominal life of 50 years. For importance class III, it recommends gI 1.3.
Bridges that are not critical for communications, or considered not economically justied to
design for the standard bridge design life of 50 years, are recommended by Eurocode 8 to be
classied in importance class I, with a recommended value gI 0.85.

Clause 2.1(3),
The reference peak ground acceleration, agR , corresponds to the reference return period, TNCR ,
Annex A [2]
of the design seismic action for bridges of ordinary importance.
Clauses 2.1(1), 2.1(3),
Note that, under the Poisson assumption of earthquake occurrence (i.e. that the number of earth2.1(4) [1]

quakes in an interval of time depends only on the length of the interval in a time-invariant way),
the return period, TR , of seismic events exceeding a certain threshold is related to the probability
that this threshold will be exceeded, P, in TL years as
TR TL/ln(1  P)

(D3.4)

So, for a given TL (e.g. the conventional design life of TL 50 years) the seismic action may
equivalently be specied either via its mean return period, TR , or its probability of exceedance
in TL years, PR .
Values of the importance factor greater or less than 1.0 correspond to mean return periods longer
or shorter, respectively, than TNCR . It is within the authority of each country to select the value of
TNCR that gives the appropriate trade-off between economy and public safety in its territory, as
well as the importance factors for bridges other than ordinary, taking into account the specic
regional features of the seismic hazard. Part 1 of Eurocode 8 recommends the value
TNCR 475 years.
The mean return period, TR(ag), of a peak ground acceleration exceeding a value ag is the inverse
of the annual rate, la(ag), of exceedance of this acceleration level:
TR(ag) 1/la(ag)

(D3.5)

A functional form commonly used for la(ag) is

la(ag) Ko(ag)  k

(D3.6)

If the exponent k (the slope of the hazard curve la(ag) in a log-log plot) is approximately
constant, two peak ground acceleration levels ag1 , ag2 , corresponding to two different mean
return periods, TR(ag1), TR(ag2), are related as


ag1
TR ag1 1=k

ag2
TR ag2
20

D3:7

Chapter 3. Seismic actions and geotechnical aspects

The value of k characterises the seismicity of the site. Regions where the difference in the peak
ground acceleration of frequent and very rare seismic excitations is very large, have low k
values (around 2). Large values of k (k . 4) are typical of regions where high ground acceleration
levels are almost as frequent as smaller ones.
Tall free-standing piers, decks built as free cantilevers or incrementally launched, etc., may be
much more vulnerable to earthquake than after completion of the full bridge. It is up to the
designer or the owner of the bridge to specify the seismic performance requirements before completion of the project and the corresponding compliance and verication criteria. Equation
(D3.4) may be used then to determine the mean return period of the seismic action that has a
given probability of being exceeded P (e.g. P 0.05) in the full duration of the bridge construction, Tc , to be used in Eq. (D3.4) in lieu of TL . This mean return period may be used as TR(ag1) in
Eq. (D3.7), to compute the peak ground acceleration, ag1 , with a probability P of been exceeded
during construction. In that case, ag2 agR and TR(ag2) TNCR .
3.1.2.3 Horizontal elastic response spectrum
The Eurocode 8 spectra include ranges of:
g
g
g

Annex A [2]

Clauses 3.2.2.1(3),
3.2.2.2(1) [1]

constant spectral pseudo-acceleration for natural periods between TB and TC


constant spectral pseudo-velocity between periods TC and TD
constant spectral displacement for periods longer than TD .

The elastic response spectral acceleration for any horizontal component of the seismic action is
described in Parts 1 and 2 of Eurocode 8 by the following expressions:
Short-period range:



T 
2:5h  1
0  T  TB : Sa T ag S 1
TB


D3:8a

Constant spectral pseudo-acceleration range:


TB  T  TC : Sa T ag S  2:5h

D3:8b

Constant spectral pseudo-velocity range:


TC  T  TD : Sa T ag S  2:5h

 
TC
T

D3:8c

Constant spectral displacement range:



TD  T  4 s: Sa T ag S  2:5h

TC TD
T2


D3:8d

where ag is the design ground acceleration on rock and S is the soil factor.
p
h 10=5 j  0:55 (Bommer and Elnashai, 1999) is a correction factor for viscous damping
ratio, j, other than the reference value of 5% (from Parts 1 and 2 of Eurocode 8); the value
j 5% is considered to be representative of cracked reinforced concrete. The viscous
damping values specied in Part 2 of Eurocode 8 for components of various structural materials
are shown in Table 3.1.

Clause 3.2.2.2(3) [1]


Clauses 4.1.3(1),
7.5.4(3) [2]

Note the uniform amplication of the entire spectrum by the soil factor, S, over the spectrum for
rock. By denition, S 1 over rock. The value agS plays the role of effective ground acceleration,
as the spectral acceleration at the constant spectral acceleration plateau is always equal to 2.5agS.
The values of the periods TB , TC and TD (i.e. the extent of the ranges of constant spectral pseudoacceleration, pseudo-velocity and displacement) and of the soil factor, S, are taken to depend
mainly on the ground type. In the Eurocodes the term ground includes any type of soil and

Clauses 3.2.2.2(2),
3.1.2(1), [1]
21

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 3.1. Values of viscous damping for different structural materials


Material

Damping: %

Reinforced concrete components


Pre-stressed concrete components
Welded steel components
Bolted steel components

5
2
2
4

rock. Parts 1 and 2 of Eurocode 8 recognise ve standard ground types, over which it recommends values for TB , TC , TD and S, and two special ones, as listed in Table 3.2.
Clauses 3.1.2(1)
3.1.2(3) [1]

The characterisation of the ground is based on the average value of the shear wave velocity, vs,30 ,
at the top 30 m (CEN, 2004a):
30
vs;30 X
hi

D3:9

v
i 1;N i
where hi and vi are the thickness (in metres) and the shear wave velocity at small shear strains (less
than 10  6) of the ith layer in N layers. If the value of vs,30 is not known, for soil types B, C or D
Part 1 of Eurocode 8 allows the use of alternatives to characterise a soil: for cohesionless soils
especially, the SPT (Standard Penetration Test) blow-count number may be used (e.g. according
to the correspondence of SPT to vs,30 in Ohta and Goto (1976)); for cohesive soils, the undrained
cohesive resistance (cu).
Clause 3.1.2(4) [1]

Clause 3.2.2.2(2) [1]

The two special ground types, S1 and S2, require the carrying out of special site-specic studies to
dene the seismic action (CEN, 2004a). For ground type S1, the study should take into account
the thickness and the vs value of the soft clay or silt layer and the difference from the underlying
materials, and should quantify their effects on the elastic response spectrum. Note that soils of
type S1 may have low internal damping and exhibit linear behaviour over a large range of
strains, producing peculiar amplication of the bedrock motion and unusual or abnormal
soilstructure interaction effects. The scope of the site-specic study should also address the
possibility of soil failure under the design seismic action (especially at ground type S2 deposits
with liqueable soils or sensitive clays) (CEN, 2004a).
The values of TB , TC , TD and S for the ve standard ground types A to E are meant to be dened
by each country in the National Annex to Eurocode 8, depending on the magnitude of earthquakes contributing most to the hazard. The geological conditions at the site may also be
taken into account in addition, to determine these values. In principle, S factors that decrease
with increasing spectral value because of the soil nonlinearity effect may be introduced.
Instead of spectral amplication factors that decrease with increasing design acceleration
Table 3.2. Ground types in Part 1 of Eurocode 8 for the denition of the seismic action

A
B
C
D
E
S1
S2

22

Description

vs,30: m/s

NSPT

cu: kPa

Rock outcrop, with less than 5 m cover of weaker material


Very dense sand or gravel, or very stiff clay, several tens of metres
deep; mechanical properties gradually increase with depth
Dense to medium-dense sand or gravel, or stiff clay, several
tens to many hundreds of metres deep
Loose-to-medium sand or gravel, or soft-to-rm clay
520 m surface alluvium layer type C or D underlain by stiffer
material (with vs . 800 m/s)
.10 m thick soft clay or silt with plasticity index .40 and high
water content
Liqueable soils; sensitive clays; any soil not of type A to E or S1

.800
360800

.50

.250

1550

70250

180360
,180

,15

,100

,70

1020

Chapter 3. Seismic actions and geotechnical aspects

Table 3.3. Recommended parameter values in Part 1 of Eurocode 8 for standard horizontal elastic
response spectra
Ground type

Spectrum type 1

A
B
C
D
E

Spectrum type 2

TB: s

TC: s

TD: s

TB: s

TC: s

TD: s

1.00
1.20
1.15
1.35
1.40

0.15
0.15
0.20
0.20
0.15

0.4
0.5
0.6
0.8
0.5

2.0
2.0
2.0
2.0
2.0

1.0
1.35
1.50
1.80
1.60

0.05
0.05
0.10
0.10
0.05

0.25
0.25
0.25
0.30
0.25

1.2
1.2
1.2
1.2
1.2

(spectral or ground) as in US codes, the non-binding recommendation of a note in Part 1 of


Eurocode 8 is for two types of spectra:
g
g

Type 1: for moderate- to large-magnitude earthquakes.


Type 2: for low-magnitude ones (e.g. with a surface magnitude less than 5.5) at close
distance, producing over soft soils motions rich in high frequencies.

The values of TB , TC , TD and S recommended in a non-binding note in Part 1 of Eurocode 8 for


the ve standard ground types A to E are given in Table 3.3 They are based on Rey et al. (2002)
and European strong motion data. There are certain regions in Europe (e.g. where the hazard is
contributed mainly by strong intermediate-depth earthquakes, as in the part of the eastern
Balkans affected by the Vrancea region) where the two recommended spectral shapes may not
be suitable. The lower S values of type 1 spectra are due to the larger soil nonlinearity in the
stronger ground motions produced by moderate to large-magnitude earthquakes. Figure 3.1
depicts the recommended type 1 spectral shape.
The values recommended in Part 1 of Eurocode 8 for the period TD at the outset of the constant
spectral displacement region seem rather low. Indeed, for exible structures, such as bridges with
tall piers or supported only on movable bearings, they may not lead to safe-sided designs.
Accordingly, Part 2 of Eurocode 8 calls the attention of designers and national authorities to
the fact that the safety of structures with seismic isolation depends mainly on the displacement

Clause 7.4.1(1) [2]


Clause 3.2.2.5(8) [1]

Figure 3.1. Elastic response spectra of type 1 recommended in Eurocode 8, for a peak ground
acceleration on rock equal to 1g and for 5% damping
4
Soil A
Soil B
Soil C
Soil D
Soil E

3.5
3

Sa /ag

2.5
2
1.5
1
0.5
0

0.5

1.5

2.5

3.5

T: s

23

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

demands on the isolating system that are directly proportional to the value of TD . So, as allowed
in Part 1 of Eurocode 8 specically for design with seismic isolation or energy dissipation devices,
Part 2 invites its National Annex to specify a value of TD for bridges with such a design that is
more conservative (longer) than the value given in the National Annex to Part 1. If the National
Annex to Part 2 does not exercise this right for national choice, the designer may do so, taking
into account the specic conditions of the seismically isolated bridge.
Clause 3.2.2.5(4) [1]

Clauses 4.1.7(1)
4.1.7(3) [2]

A safeguard against the rapid decay of the elastic spectrum for T > TD , is provided by the lower
bound of 0.2ag recommended in Part 1 of Eurocode 8 for the design spectral accelerations (see
Eqs (D3.12c) and (D3.12d) below).
3.1.2.4 Elastic spectra of the vertical component
The vertical component of the seismic action needs to be taken into account only in the seismic
design of (CEN, 2005):
g
g
g
g

Clause 3.2.2.3(1) [1]


Clause 3.2.2.2(1) [2]

prestressed concrete decks (acting upwards)


any bearings
any seismic links
piers in zones of high seismicity:
if the pier is already taxed by bending due to permanent actions on the deck or
the bridge is located within 5 km of an active seismotectonic fault (dened as one where
the average historic slip rate is at least 1 mm/year and there is topographic evidence of
seismic activity in the past 11 000 years), in which case site-specic spectra that account
for near-source effects should be used.

Eurocode 8 gives in Part 1 a detailed description of the vertical elastic response spectrum, as
follows:
Short-period range:

0  T  TB : Sa;vert T avg


T 
1
3h  1
TB


D3:10a

Constant spectral pseudo-acceleration range:


TB  T  TC : Sa;vert T avg  3h

D3:10b

Constant spectral pseudo-velocity range:


TC  T  TD : Sa;vert T avg  3h

 
TC
T

D3:10c

Constant spectral displacement range:



TD  T  4 s: Sa;vert T avg  3h

TC TD
T2


D3:10d

The main differences between the horizontal and the vertical spectra lie:
g
g

in the value of the amplication factor in the constant spectral pseudo-acceleration


plateau, which is 3 instead of 2.5
in the absence of a uniform amplication of the entire spectrum due to the type of soil.

The values of control periods TB , TC , TD are not the same as for the horizontal spectra. Part 1 of
Eurocode 8 recommends in a note the following non-binding values of TB , TC , TD and the design
ground acceleration in the vertical direction, avg:
g
g

24

TB 0.05 s
TC 0.15 s

Chapter 3. Seismic actions and geotechnical aspects

g
g
g

TD 1.0 s
avg 0.9ag , if the type 1 spectrum is considered as appropriate for the site
avg 0.45ag , if the type 2 spectrum is chosen.

The vertical response spectrum recommended in Eurocode 8 is based on work and data specic to
Europe (Ambraseys and Simpson, 1996; Elnashai and Papazoglou, 1997). The ratio avg/ag is
known to be higher at short distances (epicentral or to causative fault). However, as distance
does not enter as a parameter in the denition of the seismic action in Eurocode 8, the type of
spectrum has been chosen as the parameter determining this ratio, on the basis of the nding
that avg/ag also increases with increasing magnitude (Ambraseys and Simpson, 1996; Abrahamson
and Litehiser, 1989), which in turn determines the selection of the type of spectrum.
3.1.2.5 Topographic amplication of the elastic spectrum
Eurocode 8 provides for topographic amplication (ridge effect, etc.) of the seismic action for all
types of structures. According to Part 1 of Eurocode 8, topographic amplication of the full
elastic spectrum is mandatory for structures (including bridges) of importance above ordinary.
An informative annex in Part 5 of Eurocode 8 (CEN, 2004b) recommends amplication
factors equal to 1.2 over isolated cliffs or long ridges with a slope (to the horizontal) less than
308, or to 1.4 at ridges steeper than 308. However, as a bridge on a ridge is fairly rare, the
need to account for such an effect would be exceptional.
3.1.2.6 Near-source effects
Directivity effects of the seismic motion along the direction of rupture propagation are usually
observed near the seismotectonic fault rupture in a land strip parallel to the fault. This may show
up at the site as a large velocity pulse of long period in the direction transverse to the fault. The
general rules of Eurocode 8 in Part 1 do not provide for near-source effects. However, Part 2 of
Eurocode 8 (CEN, 2005) requires elaboration of site-specic spectra that take into account
near-source effects if the bridge is within 10 km horizontally from a known active fault that
may produce an event of moment magnitude higher than 6.5. In this respect, it gives a default
denition of an active fault as one where the average historic slip rate is at least 1 mm/year
and there is topographic evidence of seismic activity within the Holocene period (i.e. during
the past 11 000 years).

Clause 3.2.2.1(6) [1]


Annex A [3]

Clause 3.2.2.3(1) [2]

For bridges less than 15 km from a known active fault, the Caltrans Seismic Design Criteria
(Caltrans, 2006) increase spectral ordinates by 20% for all periods longer than 1 s, while
leaving them unchanged for periods shorter than 0.5 s, with linear interpolation in the period
range in between. Although not stated in Caltrans (2006), this increase, known as the directivity
effect, should only be restricted to the fault normal component of the ground motion, leaving the
fault parallel component unaffected (Sommerville et al., 1997).
It should be noted that near-source effects are quite usual in seismic-prone areas.
3.1.2.7 Design ground displacement and velocity
The value given in Part 1 of Eurocode 8 for the design ground displacement, dg , corresponding to
the design ground acceleration, ag , is based on the assumption that the spectral displacement
within the constant spectral displacement range, derived as (T/2p)2Sa(TD) with the spectral acceleration at T TD given from Eq. (D3.8c), entails an amplication factor of 2.5 over the ground
displacement that corresponds to the design ground acceleration, ag . Taking (2p)2  40, we
obtain (for ag in m/s2, not in g)
dg 0:025ag STC TD

Clause 3.2.2.4(1) [1]


Clauses 3.3(6),
6.6.4(3) [2]

D3:11

Equation (D3.11) gives estimates of the ratio dg/ag that are rather on the high side compared with
more detailed predictions as a function of magnitude and distance on the basis of Bommer and
Elnashai (1999) and Ambraseys et al. (1996).
The design ground velocity vg may be obtained from the design ground acceleration ag as follows:
vg STC ag/2p

Clause 6.7.4.(3) [2]

(D3.12)
25

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Clauses 2.1(2),
3.2.4(1), 4.1.6(1) [2]
Clause 3.2.2.5(4) [1]

3.1.3
Design spectrum for elastic analysis
For the horizontal components of the seismic action the design spectrum in Eurocode 8 is:
Short-period range:

0  T  TB : Sa;d T ag S



2 T 2:5 2


3 TB q
3

D3:13a

Constant spectral pseudo-acceleration range:


TB  T  TC : Sa;d T ag S

2:5
q

Constant spectral pseudo-velocity range:


 
2:5 TC
TC  T  TD : Sa;d T ag S
 b ag
q T
Constant spectral displacement range:


2:5 TC TD
TD  T: Sa;d T ag S
 b ag
q
T2

D3:13b

D3:13c

D3:13d

The behaviour factor, q, in Eqs (D3.13) accounts for ductility and energy dissipation, as well as
for values of damping other than the default of 5% (see also Section 2.3.2 of this Guide).
The value 2/3 in Eq. (D3.13a) is the inverse of the overstrength factor of 1.5 considered in
Eurocode 8 to always be available even without any design measures for ductility and energy
dissipation. The factor b in Eqs (D3.13c) and (D3.13d) gives a lower bound for the horizontal
design spectrum, acting as a safeguard against excessive reduction of the design forces due to
exibility of the system (real or presumed in the design). Its recommended value in Part 1 of
Eurocode 8 is 0.2. Its practical implications may be particularly important, in view of the
relatively low values recommended by Eurocode 8 for the corner period TD at the outset of
the constant spectral displacement range.
Clause 3.2.2.5(5) [1],
Clause 4.1.6(12) [2]

Clauses 3.2.3(1),
3.2.3(2) [2]
Clauses 3.2.3.1.1(1),
3.2.3.1.1(2),
3.2.3.1.2(4) [1]
Clauses 3.2.3.1.1(3),
3.2.3.1.2(3),
3.2.3.1.3(1) [1]
Clauses 3.2.3(1),
3.2.3(2), 3.2.3(4),
3.2.3(8) [2]

26

The design spectrum in the vertical direction is obtained by substituting in Eqs (D3.13) the design
ground acceleration in the vertical direction, avg , for the effective ground acceleration, agS, and
using the values in Section 3.1.2.4 for the three corner periods. There is no clear, well-known
energy dissipation mechanism for the response in the vertical direction. So, the behaviour
factor q in that direction is taken equal to 1.0.
3.1.4
Time-history representation of the seismic action
Representation of the seismic action merely by its 5%-damped elastic response spectrum is sufcient for linear or nonlinear static analysis. For a nonlinear dynamic (response-history) analysis,
time histories of the ground motion are needed, conforming on average to the 5%-damped elastic
response spectrum dening the seismic action. Eurocode 8 (Parts 1 and 2) requires as input for a
response history analysis an ensemble of at least three records, or pairs or triplets of different
records, for analysis under two or three concurrent components of the action.
Part 1 of Eurocode 8 accepts for this purpose historic, articial or simulated records, while
Part 2 mentions only historic, modied historic or simulated records.
Articial (or synthetic) records can be mathematically produced using random vibration
theory to match almost perfectly the response spectrum dening the seismic action (Gasparini
and Vanmarcke, 1976). It is fairly straightforward to adjust the phases of the various sinusoidal
components of the articial waveform, as well as the time evolution of their amplitudes
(envelope function), so that the articial record resembles a specic recorded motion. This is
the modied historic type of record mentioned in Part 2 of Eurocode 8. Figure 3.2 shows an
example of such a record and Figure 8.47 another one. Note, however, that records that are
equally rich in all frequencies are not realistic. Moreover, an excitation with a smooth

Chapter 3. Seismic actions and geotechnical aspects

Figure 3.2. Herzegnovi X record from the 1979 Montenegro earthquake modulated to match the
Eurocode 8 type 1 spectrum for ground type C with a peak ground acceleration of 0.1g
1

Modified Hercegnovi

0.6

Target spectrum
Modified Hercegnovi

0.8

2.5
2

0.2

Sa: m/s2

a: m/s2

0.4
0

1.5

0.2

0.4
0.6

0.5

0.8
1

6
Time: s

10

12

0.5

1.5
2
Period, T: s

2.5

response spectrum without peaks or troughs introduces a conservative bias in the response, as it
does not let the inelastic response help the structure escape from a spectral peak to a trough at a
longer period. Therefore, historic records are favoured in Part 2 of Eurocode 8.
Records simulated from mathematical source models, including rupture, propagation of the
motion through the bedrock to the site and, nally, through the subsoil to the surface are also
preferred over articial ones, as the nal record resembles a natural one and is physically
appealing. Obviously, an equally good average tting of the target spectrum requires more
appropriately selected historic or simulated records than articial ones. Individual
recorded or simulated records should, according to Part 1 of Eurocode 8, be adequately qualied
with regard to the seismogenetic features of the sources and to the soil conditions appropriate to
the site (Part 1 of Eurocode 8). In plainer language, they should come from events with the
magnitude, fault distance and mechanism of rupture at the source consistent with those of the
design seismic action (Part 2 of Eurocode 8). The travel path and the subsoil conditions should
preferably resemble those of the site. These requirements are not only hard to meet but may
also conict with conformity (in the mean) to the target spectrum of the design seismic action.
The requirement in Part 1 of Eurocode 8 to scale individual historic or simulated records so that
their peak ground acceleration (PGA) matches on average the value of agS of the design seismic
action may also be considered against physical reality. It is more meaningful, instead, to use individual historic or simulated records with PGA values already conforming to the target value of agS.
Note also that the PGA alone may be articially increased or reduced, without affecting at all the
structural response. So, it is more meaningful to select the records on the basis of conformity of
spectral values alone along the lines of Part 2 of Eurocode 8 (CEN, 2005), as described below.
If pairs or triplets of different records are used as the input for analysis under two or three
concurrent seismic action components, conformity to the target 5%-damped elastic response
spectrum may be achieved by scaling the amplitude of the individual records as follows (see
Figures 8.48 and 8.49 for an application example):
g
g

Clauses 3.2.3(3),
3.2.3(6), 3.2.3(7) [2]

For each earthquake consisting of a triplet of translational components, the records of


horizontal components are checked for conformity separately from the vertical one.
The records of the vertical component, if considered, are scaled so that the average 5%damped elastic spectrum of their ensemble is at least 90% of the 5%-damped vertical
spectrum at all periods between 0.2Tv and 1.5Tv , where Tv is the period of the lowest
mode having a participation factor of the vertical component higher than those of both
horizontal ones.
For analysis in 3D under both horizontal components, the 5%-damped elastic spectra of
the two horizontal components in each pair are combined by applying the SRSS rule at
each period value. The average of the SRSS spectra of the two horizontal components of
27

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

p
the individual earthquakes in the ensemble should be at least 0.9 2  1.3 times the target
5%-damped horizontal elastic spectrum at all periods from 0.2T1 up to 1.5T1 , where T1 is
the lowest natural period of the structure (the effective period of the isolation system in
seismically isolated bridges) in any horizontal direction. If this is not the case, all
individual horizontal components are scaled up, so that their nal average SRSS
spectrum exceeds by a factor of 1.3 the target 5%-damped horizontal elastic spectrum
everywhere between 0.2T1 and 1.5T1 .
Clause 3.2.3.1.2(4) [1] Note in this respect that for analysis under a single horizontal component, Part 1 of Eurocode 8

requires the mean 5%-damped elastic spectrum of the applied motions to not fall below 90% of
that of the design seismic action at any period from 0.2T1 to 2T1 .
Clause 4.2.4.3(1) [2]
If the response is obtained from at least seven nonlinear time-history analyses with (triplets or
Clause 4.3.3.4.3(3) [1] pairs of ) ground motions chosen in accordance with the previous paragraphs, the relevant veri-

cations may use the average of the response quantities from all these analyses as the action
effect. Otherwise, it should use the most unfavourable value of the response quantity among
the (three to six) analyses.

Clauses 3.3(1)3.3(8),
Annex D [2]

3.1.5
Spatial variability of the seismic action
Unlike typical buildings, bridges are extended structures. Therefore, it is very likely that the foundation supports experience different ground motions owing to the spatial variability of the
seismic motion. This phenomenon is known by the generic term decorrelation. Decorrelation
of seismic motions arises from three different causes:
The travelling wave effect: except for vertically propagating waves, the seismic waves
exhibit an apparent velocity in the horizontal direction, causing out-of-phase motions
along the bridge, even when the amplitude remains the same.
2 Scattering of waves, especially at high frequencies: waves travelling through a
heterogeneous soil medium are scattered (diffracted and/or reected) at every
heterogeneity, no matter whether a small lens or an abrupt change in the mechanical
characteristics of the media. As the frequency increases, the wavelength decreases and the
waves perceive and are affected by smaller heterogeneities. Scattering causes the motion at
two adjacent locations to be different.
3 Propagation through different soil proles under each pier location: if the bridge
foundations are not very close to each other, the soil prole under them may be different,
causing different soil amplication from the bedrock.
1

Effects 1 and 3 can, theoretically, be accounted for (although under simplifying assumptions), if the
direction of propagation of the seismic waves is given and the soil proles are accurately known.
Effect 2 does not lend itself to calculation, without complete knowledge of the subsoil conditions
between the seismic source and the bridge. Effect 1 can easily be modelled, by shifting the ground
motion time histories by a time lag equal to the distance between the piers divided by the
apparent travelling wave velocity. Analytical studies and observations suggest that the apparent
wave velocity is not equal to the wave velocity in the upper soil layers, but rather to the wave
velocity at a signicant depth (in the rock medium where the rupture initiates); typical values
exceed 1000 m/s. Effect 3 can be computed through one-dimensional site response analysis. For
effect 2, only random vibration models with empirically determined parameters can be used.
Informative Annex D in Part 2 of Eurocode 8 presents the theory for the generation of incoherent
ground motions, including all three effects above, as well as the mathematical tools for the
analysis of the bridge response under multi-support excitations. However, because the theory
is complex, requiring specic tools for its numerical implementation and statistical site data
for the determination of model parameters, the code allows the use of a simplied approach
to take into account the spatial variability of seismic motions along the bridge. This spatial variability should be taken into account whenever the soil properties along the bridge vary and the
ground type according to Table 3.2 differs from one pier to another, or if the length of a continuous deck exceeds Lg/1.5, where Lg is the correlation distance (beyond which the motion may be
assumed to be fully uncorrelated). Recommended values of the correlation distance are given in
Table 3.4.
28

Chapter 3. Seismic actions and geotechnical aspects

Table 3.4. Values recommended in Part 5 of Eurocode 8 for the distance beyond which ground motions
may be considered as uncorrelated
Ground type

Lg : m

600

500

400

300

500

The simplied methodology consists of combining via the SRSS (square root of the sums of the
square) rule the dynamic effects of a uniform ground motion acting at every foundation, to the
effects of differential displacements imposed statically at each foundation point. Two patterns
are used for the static imposed displacements, and the results of the most unfavourable one
are retained:
g

a pattern with foundation displacements all in the same direction and proportional to the
distance along the bridge and to the ground displacement, but inversely proportional to
the correlation distance, combined with a small offset at any intermediate pier
a pattern with displacements alternating between consecutive piers.

The same patterns of imposed displacements, with an increased safety factor, are also used in the
checks of deck unseating at movable joints (see Eq. (D6.34) in Sect. 6.8.1.2).
Unlike certain truly dynamic approaches, the simplied method above cannot capture dynamic
features of the spatial variability of the seismic action, as it is essentially a pseudo-static addition
of imposed support displacements. It accounts, however, to a certain degree of approximation, for all three main effects of the spatial variability (Sextos et al., 2006; Sextos and
Kappos, 2009).

3.2.

Siting and foundation soils

3.2.1
Introduction
Signicant damage to foundations may be caused by soil-related phenomena: fault rupture, slope
instability near the bridge, liquefaction or densication of the soil due to the ground shaking.
Except in very few cases, such adverse effects cannot be accommodated by a foundation
design. Therefore, these ground hazards should be thoroughly investigated and properly mitigated to the largest possible extent.

Clause 4.1.1(1) [3]

3.2.2
Seismically active faults
Seismological evidence suggests that, where seismogenic activity is conned in the upper 20 km or Clause 4.1.2(1), (2) [3]
so of the Earths crust, co-seismic surface rupture tends to occur only if the earthquake has a Clause 3.2.2.3(1) [2]
moment magnitude Mw over about 6.5. Therefore, in Europe, surface faulting is a rather rare
event, except in Turkey and maybe in Greece or Italy. As pointed out in Part 5 of Eurocode 8,
assessment of the surface fault rupture hazard at a site requires special geological investigations,
to show that there is no active fault nearby. Ofcial documents published by competent national
authorities may, of course, map the seismically active faults. Note, though, that there are no
absolute criteria to characterise a fault as seismically active and to consider a site as close to
it. It is suggested in Part 5 of Eurocode 8 that evidence of movement in the late Quaternary
period (10 000 years) or lack of it is used as a criterion, while Part 2 of Eurocode 8 denes an
active fault as one where the average historic slip rate is at least 1 mm/year and there is
topographic evidence of seismic activity in the Holocene period (i.e. in the past 11 000 years).
A distance of several tens of metres may be used as the criterion for the immediate vicinity to
a fault.
3.2.3
Slope stability
When a structure is to be built near a natural or man-made slope, a verication of the slope stability under the seismic action shall be carried out. Although the stability checks recommended in
Part 5 of Eurocode 8 aim at ensuring a prescribed safety factor, the underlying criterion is an
ultimate limit state (ULS) beyond which the permanent displacements it entails become unacceptable. This is reected in the method of analysis proposed in Part 5 of Eurocode 8 for achieving a
safety factor above 1.0, notably a pseudo-static approach with seismic forces taken as equal to

Clauses 4.1.3.1(1),
4.1.3.1(2), 4.1.3.3(1),
4.1.3.3(3)4.1.3.3(6),
4.1.3.3(8), 4.1.3.4(4),
Annex A [3]

29

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

the product of the potential sliding mass multiplied by 50% of the design PGA at the soil surface,
agS, including the topographic amplication factor of Section 3.1.2.5, if relevant. The 50%
fraction has been chosen empirically and with back-analysis of observed performance of
slopes in earthquakes. It nevertheless reects the observation, rst pointed out by Newmark in
his 1965 Rankine lecture, that when the maximum seismic action, equal to the product of the
potentially sliding mass multiplied by the PGA, is slightly exceeded, only permanent displacements occur without a catastrophic failure. With the recommended value of the ground acceleration, 0.5agS, it is expected that the induced displacements will not exceed a few centimetres.
The design seismic inertia forces in the horizontal and vertical direction are given by
FH 0.5agSM
FV +0.5FH
FV +0.33FH

(D3.14)
if avg . 0.6ag
if avg  0.6ag

(D3.15a)
(D3.15b)

where M is the potentially sliding mass and agS includes the topographic amplication factor of
Section 3.1.2.5, if relevant. The seismic design resistance of the soil should be calculated with the
soil strength parameters divided by the appropriate partial factor, dened in the National Annex
to Part 5 of Eurocode 8.
It is essential not to overlook the applicability conditions of the pseudo-static method of analysis,
that is:
g
g

The geometry of the topographic prole and the ground prole are reasonably regular.
The ground materials of the slope and the foundation, if water saturated, are not prone to
developing signicant pore water pressure build-up that may lead to loss of shear strength
and stiffness degradation under seismic conditions. The same limitation applies to certain
unusual soils, such as sensitive clays, although the mechanism of strength degradation is
different.

For high values of agS it may prove hard to verify the slope stability using the pseudo-static
method. If so, the designer may opt for computing the actual induced permanent displacements
and checking whether they are acceptable. A simplied way to estimate the displacements is the
Newmark sliding block method. This entails the preliminary choice of the most critical sliding
surface and the associated critical value of agS for which the safety factor drops to 1.0. With
the selection of appropriate time histories for the ground motion, double integration of the difference of the input acceleration and the critical one is carried out over the time intervals during
which the former exceeds the latter. The outcome may be taken as the permanent slope displacement along the chord of the critical circular failure surface. More rened analyses, accounting for
the seismic response of the slope in the evaluation of the rigid block acceleration, may be
warranted in some cases.
One essential requirement for the application of pseudo-static analysis or of the Newmark
sliding block method is that the soil strength does not vary signicantly during the earthquake.
When the strength is reduced by a pore pressure build-up, it may be evaluated through the
expression


Du
tan wr 1  0 tan w
sv

D3:16

where wr is the reduced friction angle, Du the pore pressure increase estimated from empirical
correlations or preferably from experimental tests, and s v0 is the effective vertical stress.

3.3.

Soil properties and parameters

3.3.1
Introduction: the meaning of soil property values
Many geotechnical tests, particularly eld tests, do not allow determining directly the value of
basic geotechnical parameters or coefcients, notably for strength and deformations. Instead,
30

Chapter 3. Seismic actions and geotechnical aspects

these values are to be derived via theoretical or empirical correlations. Part 2 of Eurocode 7
(CEN, 2007) denes derived values as:
Derived values of geotechnical parameters and/or coefcients are obtained from test results by
theory, correlation or empiricism. Derived values of a geotechnical parameter then serve as
input for assessing the characteristic value of this parameter in the sense of Eurocode 7 Part
1 and, further, its design value, by applying the partial factor gM (material factor).
The philosophy regarding the denition of characteristic values of geotechnical parameters is
given in Part 1 of Eurocode 7 (CEN, 2003):
The characteristic value of a geotechnical parameter shall be selected as a cautious estimate of
the value affecting the occurrence of the limit state . . . the governing parameter is often the
mean of a range of values covering a large surface or volume of the ground. The characteristic
value should be a cautious estimate of this mean value.
These excerpts from Eurocode 7 reect the concern that we should be able to keep using the
values of the geotechnical parameters traditionally used, whose determination is not standardised
(they often depend on the judgment of the geotechnical engineer). However, two remarks are due
in this connection: on one hand, the concept of a derived value of a geotechnical parameter (preceding the determination of the characteristic value) has been introduced, but, on the other, there
is now a clear reference to the limit state involved and the assessment of a spatial mean value, as
opposed to a local value; this might appear as a specic feature of geotechnical design which,
indeed, involves large surface areas or large ground volumes.
Statistical methods are mentioned in Eurocode 7 only as a possibility:
If statistical methods are used, the characteristic value should be derived such that the
calculated probability of a worse value governing the occurrence of the limit state under
consideration is not greater than 5%.
The general meaning is that the characteristic value of a geotechnical parameter should not be
very different from the values traditionally used. Indeed, for the majority of projects, the geotechnical investigation is such that no meaningful statistical treatment of the data can be performed.
Statistical methods are, of course, useful for very large projects where the amount of data justies
their use.
3.3.2
Soil properties
Eurocode 8 considers both the strength properties and the deformation characteristics. It further
recognises that earthquake loading is essentially of short duration. Consequently, most soils
behave in an undrained manner. In addition, for some of them the properties may be affected
by the rate of loading.

Clauses 3.1(1)3.1(3)
[3]

3.3.2.1 Strength parameters


For cohesive soils, the relevant strength property is the undrained shear strength, cu . For most of
them this value can be taken as equal to the conventional static shear strength. However, on the
one hand some plastic clays may be subject to cyclic degradation of strength, but, on the other,
some clays may exhibit a shear strength increase with the rate of loading. These phenomena
should ideally be given due consideration in the choice of the relevant undrained shear
strength. The recommended partial factor gM on cu is equal to 1.4.
For cohesionless soils, relevant strength properties are the drained friction angle, w0 , and the
drained cohesion, c0 . They are directly usable for dry or partially saturated soil. For saturated
soils, they require knowledge of the pore water pressure variation during cyclic loading, u,
which directly governs the shear strength through the MohrCoulomb failure criterion:

t (s  u) tan w0 c0

(D3.17)

The evaluation of u is very difcult. Therefore, Part 5 of Eurocode 8 gives an alternative approach,
namely using the undrained shear strength under cyclic loading, tcy,u , which may be determined
31

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

from experimental correlations with, for instance, the soil relative density or any other index
parameter, such as the blow counts number, N, in a standard penetration test (SPT).
The recommended values for the partial factors gM are:
g
g

Clauses 3.2(1)3.2(4),
4.2.3(1)4.2.3(3) [3]

gM 1.25 on tan(w0 ) and tcy,u


gM 1.4 on c0 .

3.3.2.2 Stiffness and damping parameters


The soil stiffness is dened by the shear wave velocity, vs , or equivalently the soil shear modulus
G. The main role played by this parameter is in the classication of the soil prole according to
the ground types in Table 3.2 of Section 3.1.2.3 in this Guide. Additional applications that
require knowledge of the shear stiffness of the soil prole include the evaluation of:
g
g

soilstructure interaction
site response analyses to dene the ground surface response for special soil categories
(prole S1).

In the applications listed above, it is essential to recognise that soils are highly nonlinear materials
and that the relevant values to use in calculations are not the elastic ones but secant values
compatible with the average strain level induced by the earthquake, typically of the order of
5  10  4 to 10  3. Part 5 of Eurocode 8 proposes the set of values in Table 3.5, depending on
the peak ground surface acceleration. Note that the fundamental variable governing the
reduction factor is the shear strain and not the peak ground surface acceleration. However, in
order to provide useful guidance to designers, the induced strains have been correlated to PGAs.
In addition to the stiffness parameters, soil internal damping should be taken into account in
soilstructure interaction analyses. The soil damping ratio also depends on the average
induced shear strain, and is correlated to the reduction factor for the stiffness, as listed in
Table 3.5.

3.4.
Clauses 4.1.4(1)
4.1.4(3) [3]

Liquefaction, lateral spreading and related phenomena

3.4.1
Nature and consequences of the phenomena
Liquefaction is a process in which cohesionless or granular sediments below the water table temporarily lose strength and behave as a viscous liquid rather than a solid, during strong ground
shaking. Saturated, poorly graded, loose, granular deposits with low nes content are most susceptible to liquefaction. Liquefaction does not occur randomly: it is restricted to certain geological and hydrological environments, primarily recently deposited sands and silts in areas
with a high ground water level. Dense and more clayey soils, including well-compacted lls,
have low susceptibility to liquefaction.
The liquefaction process itself may not necessarily be particularly damaging or hazardous. For
engineering purposes, it is not the occurrence of liquefaction that is of importance but the potential of the process and of associated hazards to damage structures. The adverse effects of liquefaction can be summarised as follows:
g

Flow failures, when completely liqueed soil or blocks of intact material ride on a layer of
liqueed soil. Flows can be large and develop on moderate to steep slopes.

Table 3.5. Average soil damping ratios and reduction factors (+1 standard deviation) for the shear wave
velocity vs and the shear modulus G within the upper 20 m of soil in Part 5 of Eurocode 8 (for soil with
vs , 360 m/s)

32

Ground acceleration, agS: g

Damping ratio: %

vs/vs,max

Gs/Gs,max

0.1
0.2
0.3

3
6
10

0.9 (+0.07)
0.7 (+0.15)
0.6 (+0.15)

0.8 (+0.1)
0.5 (+0.2)
0.36 (+0.2)

Chapter 3. Seismic actions and geotechnical aspects

g
g

g
g

Lateral spreading, with lateral displacement of supercial blocks of soil as a result of the
liquefaction of a subsurface layer. Spreading generally develops on gentle slopes, and
moves towards a free face, such as an incised river channel or coastline. It may also occur
through the failure of shallow liqueed layers subjected to a high vertical load on part of
the ground surface due to a natural or articial embankment or cut.
Ground oscillation: where the ground is at or the slope too gentle to allow lateral
displacement, liquefaction at depth may disconnect overlying soils from the underlying
ground, allowing the upper soil to oscillate back and forth in the form of ground waves.
These oscillations are usually accompanied by ground ssures and the fracture of rigid
extended structures, such as pavements and pipelines.
Loss or reduction in bearing capacity, when earthquake shaking increases pore water
pressures, which in turn cause the soil to lose its strength and bearing capacity.
Soil settlement, as the pore water pressures dissipate and the soil densies after
liquefaction. Settlement of structures may occur, owing to the reduction in the bearing
capacity or the ground displacements noted above. In piled foundations the postearthquake settlement of the liqueed layer due to pore pressure dissipation induces
negative skin friction along the shaft, in all layers above the liqueed layer.
Increased lateral pressures on retaining walls, when the soil behind a wall liquees and
behaves like a heavy uid with no internal friction.
Flotation of buried structures, when buried structures, such as tanks and pipes, become
buoyant in the liqueed soil.

Other manifestations of liquefaction, such as sand boils, can also occur, and may pose a risk to
structures, particularly through loss or reduction in the bearing capacity and settlement.
Liquefaction has been extensively studied since 1964. The state of the art is now well established
and, more importantly, allows reliable prediction of the occurrence of liquefaction. So, this
aspect is fully covered in Part 5 of Eurocode 8, including a normative annex for the use of
SPT measurements for the evaluation of the undrained cyclic strength of cohesionless soils.
However, in addition to the SPT, other techniques are allowed for the determination of the
soil strength, such as cone penetration tests (CPTs) and shear wave velocity measurements. Laboratory tests are not recommended, because to obtain a reliable estimate of liquefaction resistance, very specialised drilling and sampling techniques are needed, beyond the budget of any
common project. It should, however, be noted that there have been numerous developments
in liquefaction assessment methodologies in recent years (e.g. Seed et al., 2003; Idriss and
Boulanger, 2008). So, the methods described in Part 5 of Eurocode 8 may be potentially unconservative, especially for materials with a high nes content. It is therefore recommended that an
expert is involved in the liquefaction assessment.
3.4.2
Liquefaction assessment
Clauses 4.1.4(3)
The verication of the liquefaction susceptibility is carried out under free eld conditions, but
4.1.4(6), 4.1.4(10),
with the prevailing situation during the lifetime of the structure. For instance, if a tall
platform is going to be built to prevent ooding of the site, or the water table will be lowered 4.1.4(11), Annex B [3]
on a long-term basis, these developments should be reected in the evaluation.
The recommended analysis is a total stress analysis in which the seismic demand, represented
by the earthquake-induced stresses, is compared with the seismic capacity (i.e. the undrained
cyclic shear strength of the soil also called liquefaction resistance). The shear stress demand
is expressed in terms of a cyclic stress ratio (CSR), and the capacity in terms of a cyclic
resistance ratio (CRR). In both ratios the normalisation is with respect to the vertical effective
stress, s v0 .
A soil should be considered susceptible to liquefaction whenever CRR , l CSR, where l is a
factor of safety, with a recommended value of 1.25. The CSR is evaluated with a simplied
version of the SeedIdriss formula, which allows a rapid calculation of the induced stress
along the depth, without resorting to a dynamic site response analysis:
CSR 0:65ag S=gsv =s 0v

D3:18
33

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

where sv and s v0 are the overburden pressure and the vertical effective stress at the depth of
interest. Equation (D3.18) may not be applied for depths larger than 20 m. The liquefaction
resistance ratio, CRR, may be estimated through empirical correlations with an index parameter,
such as the SPT blow count, the static CPT point resistance or the shear wave velocity. Note that
all of these methods should be implemented with several corrections of the measured index
parameter for the effects of the overburden at the depth of measurement, the nes content of
the soil, the effective energy delivered to the rods in SPTs, etc. In normative Annex B in Part 5
of Eurocode 8, the CRR is assessed based on a corrected SPT blowcount, using empirical
liquefaction charts relating CRR t/s v0 to the corrected SPT blow count N1(60) for an earthquake surface magnitude of 7.5. Correction factors are provided in Annex B [3] for other
magnitudes.
Clause 4.1.4(7),
4.1.4(8) [3]

A soil may be prone to liquefaction if it presents certain characteristics that govern its strength
and the seismic demand is large enough. Taking the opposite view, Part 5 of Eurocode 8 introduces cumulative conditions under which the soil may be considered as not prone to liquefaction
and liquefaction assessment is not required:
g

Low ground surface acceleration, agS , 0.15g, and


soils with a clay content higher than 20% and a plasticity index above 10% or
soils with a silt content higher than 35% and a corrected SPT blow count of over 20 or
clean sands with a corrected SPT blow count higher than 30.

The assessment of liquefaction is also not required for layers located deeper than 15 m below the
foundation. This does not mean that those layers are not prone to liquefaction, although susceptibility to liquefaction decreases with depth; it means, instead, that because of their depth their
liquefaction will not affect the structure. Although it is not spelled out in Eurocode 8, obviously
this condition is not sufcient by itself: it should be complemented with a condition on the foundation dimensions relative to the layer depth.
Clauses 4.1.4(12)
4.1.4(14) [3]

If soils are found to be susceptible to liquefaction, mitigation measures, such as ground improvement and piling (to transfer loads to layers not susceptible to liquefaction), should be considered
to ensure foundation stability. The use of pile foundations alone should be considered with
caution, in view of:
g
g
g

Clauses 4.1.4(12)
4.1.4(14) [3]

the large forces induced in the piles by the loss of soil support in the liqueable layers
the post-earthquake settlements of the liqueed layers causing negative skin friction in all
layers located above
the inevitable uncertainties in determining the location and thickness of such layers.

3.4.3
Liquefaction mitigation
If soils are found to be susceptible to liquefaction and the consequences are considered
unacceptable for the structure (excessive settlement, loss of bearing capacity, etc.), ground
improvement should be considered. Several techniques are available to improve the resistance
to liquefaction: soil densication, soil replacement, sand compaction piles, drainage (gravel
drains), deep soil mixing (limecement mixing), stone columns, blasting, jet grouting, etc. The
most commonly used among these are stone columns and soil densication, the latter through
vibrootation, dynamic compaction or compaction grouting. Some techniques, such as stone
columns, offer the advantage of combining several effects, for example, densication and
drainage.
Not all techniques are appropriate for any soil condition. The most appropriate ones should be
chosen taking into account the depth to be treated, the nes content of the soil and the presence
of adjacent structures. Attention should also be paid to the efciency, durability and cost of the
solution. For instance, dynamic compaction is better suited for shallow clean sand layers; jet
grouting and stone columns may be used to improve the soil and offer a good load-bearing
layer under shallow foundations; compaction grouting, albeit more costly, is very efcient for
almost any soil, etc. Methods based on drainage should be considered with care, as drainage
conditions may change in time and clogging may occur, especially in environments with large
uctuation of the water table.

34

Chapter 3. Seismic actions and geotechnical aspects

Figure 3.3. Assessment of volumetric strain


Cyclic shear strain, cyc: %
3

Volumetric strain due to compaction, c: %

10
103

102

N1 < 40
<30
<20
<15
<10
<5

102

101

15 cycles

101

10

There is signicant experience with all the methods listed above: when properly implemented,
they have a good track record during earthquakes.
3.4.4
Settlements
The magnitude of the settlement induced by the earthquake should be assessed when there are
extended layers or thick lenses of loose, unsaturated cohesionless materials at shallow depths.
Excessive settlements may also occur in very soft clays because of cyclic degradation of their
shear strength under ground shaking of long duration. If the settlements caused by densication
or cyclic degradation appear capable of affecting the stability of the foundations, consideration
should be given to ground improvement methods.

Clauses 4.1.5(1)
4.1.5(4) [3]

Earthquake-induced settlement can be estimated using empirical relationships between volumetric strain, SPT N values (corrected for overburden) and the factor of safety against liquefaction. For example, the approach in Tokimatsu and Seed (1987) is based on relationships between
the volumetric strain, the cyclic shear strain and SPT N values. The peak shear strain computed
from the one-dimensional response analysis and the SPT-corrected N value at that point are
entered into the Tokimatsu and Seed chart (Figure 3.3) to yield the volumetric strain. The
total settlement can then be obtained by integrating these strains over depth.
3.4.5
Lateral spreading
Lateral spreading is a highly unpredictable phenomenon. Nevertheless, empirical correlations
have been developed to estimate the lateral displacement of the ground, DH, in metres due to
liquefaction (Youd et al., 2002):
log DH 16:713 1:532M  1:406 log R  0:012R 0:592 log W 0:540 log T15




3:413 log 100  F15  0:795 log D5015 0:1 mm
D3:19
where M is the moment magnitude of the earthquake; R is the nearest horizontal or map
distance from the site to the seismic energy source (if R , 0.5 km, use R 0.5 km);
R R 10(0.89M  5.64); T15 is the cumulative thickness of saturated granular layers with a
corrected SPT blow count (N1)60 less than 15; F15 is the average nes content of granular
material (fraction passing the #200 sieve) within T15; D5015 is the average mean grain size of
granular material in T15, and W is the free-face ratio, dened as the height, H, of the free face
as a percentage of the distance, L, from the base of the free face to the point in question.
35

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Such relationships should be used with care and by experienced engineers, as there is no physical
theory as yet to conrm them. Note also that large displacements, over 6 m, should not be
taken to apply in engineering practice, because they are poorly supported in the database
(Youd et al., 2002).
REFERENCES

Abrahamson NA and Litehiser JJ (1989) Attenuation of peak vertical acceleration. Bulletin of the
Seismological Society of America 79: 549580.
Ambraseys N, Simpson K and Bommer JJ (1996) Prediction of horizontal response spectra in
Europe. Journal of Earthquake Engineering and Structural Dynamics 25(4): 371400.
Ambraseys NN and Simpson KA (1996) Prediction of vertical response spectra in Europe.
Earthquake Engineering and Structural Dynamics 25(4): 401412.
Bommer JJ and Elnashai AS (1999) Displacement spectra for seismic design. Journal of Earthquake
Engineering 3(1): 132.
Caltrans (2006) Seismic Design Criteria, version 1.4. California Department of Transportation,
Sacramento, CA.
CEN (Comite Europeen de Normalisation) (2003) EN 19971:2003: Eurocode 7 Geotechnical
design Part 1: General rules. CEN, Brussels.
CEN (2004a) EN 1998-1:2004: Eurocode 8 Design of structures for earthquake resistance Part 1:
General rules, seismic actions and rules for buildings. CEN, Brussels.
CEN (2004b) EN 1998-5:2004: Eurocode 8 Design of structures for earthquake resistance Part 5:
Foundations, retaining structures, geotechnical aspects. CEN, Brussels.
CEN (2005) EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance Part 2:
Bridges. CEN, Brussels.
CEN (2007) EN 19972:2007: Eurocode 7 Geotechnical design Part 2: Ground investigation
and testing. CEN, Brussels.
Elnashai AS and Papazoglou AJ (1997) Procedure and spectra for analysis of RC structures
subjected to strong vertical earthquake loads. Journal of Earthquake Engineering 1(1): 121155.
Gasparini DA and Vanmarcke EH (1976) Simulated Earthquake Motions Compatible with
Prescribed Response Spectra. Department of Civil Engineering, Massachusetts Institute of
Technology, Cambridge, MA. Research Report R76-4.
Idriss IM and Boulanger RW (2008) Soil Liquefaction During Earthquakes. Earthquake
Engineering Research Institute, Oakland, CA. MNO-12.
Ohta Y and Goto N (1976) Estimation of S-wave velocity in terms of characteristic indices of soil.
Butsuri-Tanko 29(4): 3441.
Rey J, Faccioli E and Bommer JJ (2002) Derivation of design soil coefcients (S) and response
spectral shapes for Eurocode 8 using the European Strong-Motion Database. Journal of
Seismology 6(4): 547555.
Seed HB, Cetin KO, Moss RES et al. (2003) Recent advances in soil liquefaction engineering: a
unied and consistent framework. 26th Annual ASCE LA Geotechnical Seminar, HMS Queen
Mary, Long Beach, CA, Keynote Presentation.
Sextos AG and Kappos AJ (2009) Evaluation of seismic response of bridges under asynchronous
excitation and comparison with Eurocode 8-2 provisions. Bulletin of Earthquake Engineering 7:
519545.
Sextos AG, Kappos AJ and Kolias B (2006) Computing a reasonable spatially variable
earthquake input for extended bridge structures. First European Conference on Earthquake
Engineering and Seismology, Geneva, paper 1601.
Sommerville PG, Smith NF, Graves RW and Abrahamson NA (1997) Modication of empirical
strong ground motion attenuation relations to include the amplitude and duration effects of
rupture directivity. Seismological Research Letters 68: 199222.
Tokimatsu K and Seed HB (1987) Evaluation of settlements in sand due to earthquake shaking.
Journal of Geotechnical Engineering of the ASCE 113(8): 861878.
Youd TL, Hansen CM and Bartlett SF (2002) Revised multilinear regression equations for
prediction of lateral spread displacement. Journal of the Geotechnical and Geoenvironmental
Engineering 128(12): 10071017.

36

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.037

Chapter 4

Conceptual design of bridges for


earthquake resistance
4.1.

Introduction

During the conceptual design phase, the layout of the structure is established, the structural
materials for its various parts are selected and the construction technique and procedure are
decided. Conceptual design concludes with a preliminary sizing of all members, in order to
allow the next phase of the design process to be carried out, namely the analysis for the
calculation of the effects of the design actions (including seismic) in terms of internal forces
and deformations in structural members. Analysis is followed by the detailed design phase
(notably the verication of member sizes, the dimensioning of the reinforcement, etc., on the
basis of the calculated action effects) and the preparation of material specications, construction
drawings and any other information that is necessary or helpful for the implementation of the
design.
Conceptual design is of utmost importance for the economy, safety and tness for use of the
structure. In addition to technical skills and knowledge, it requires judgement, experience and a
certain intuition. Although conceptual design cannot be taught, several authoritative documents
(recently, b, 2009, 2012) give general principles and guidance. Useful general sources for bridges
are b (2000) and (2004), and, for their seismic design, Priestley et al. (1996) and b (2007).
The conceptual design of a bridge is controlled mainly by that of the deck, which in turn is
governed by its use, the preferred construction technique (Table 4.1), aesthetics, topography
and of course cost issues. The three last considerations signicantly inuence the conceptual
design of the piers as well. Gravity loads and the construction technique control the design of the
deck. The deck spans and their erection, alongside the terrain, determine the number and
location of the piers. In seismic regions, the piers themselves and their connection to the deck
are governed by the seismic action. In addition, seismic considerations are taken into account
for some aspects of the deck design, notably its continuity across spans and sometimes its connection with the abutments and the piers. Before delving into these purely seismic aspects, it is worth
recalling that one of the prime objectives of the conceptual design of a bridge for non-seismic
loads is to reduce the deck dead load, as this is normally the prime contributor to the action
effects in the deck, the piers and the foundation for the combination of factored gravity loads
(the persistent and transient loads combination in Eurocode terminology). This is rst pursued
through the choice of the material(s) and the shape of the deck section; once these are chosen,
an effort is made to reduce its dimensions by using higher-strength materials, external prestressing (if relevant), etc. Needless to say, although the conceptual and detailed seismic design of
bridges concerns mainly the piers and the way they are connected to or support the deck,
reducing the decks self-weight is of prime importance for the bridge seismic design as well:
both seismic force and displacement demands increase with the deck mass (they are normally
approximately proportional to its square root), and there is good reason to reduce it.

Clauses 2.4(1)2.4(3)
[2]

As pointed out in Section 2.3.1 of this Guide, the prime decision in the conceptual seismic design
of the bridge is how to accommodate the horizontal seismic displacements of the deck with
respect to the ground. Four options were highlighted there, repeated below for completeness:
1

to support the deck on all abutments and piers through bearings (or similar devices) that
can slide or are horizontally very exible
37

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 4.1. Span range, erection speed and continuity of the deck across spans and with the piers for
different erection techniques
Deck erection

Normal span
range: m

Erection speed:
m/week

Full deck
continuity

Integral deck
and piers

Prefabricated girders
On scaffolding/falsework on grade
On mobile launching/casting girder
or gantry (span-by-span)
Free/balanced cantilever
Cast-in-situ deck segments
Prefabricated deck segments
Incremental launching
Without temporary props
With temporary props

1050
550
3060

25100a
510
1050

Normally not
Normally yes
Yes

Normally not
Yes or no
Yes or no

60300
40160

615
2060a

Yes

Normally yes

3070
70120

1030
1030

Yes

Normally not

Speed depends on the capacity of the prefabrication plant

to x the deck to the top of at least one pier (but not at the abutments) and let those piers
accommodate the horizontal seismic displacements through inelastic rotations in exural
plastic hinges
3 to let the base of the piers slide with respect to the soil or allow inelastic deformations to
develop in foundation piles
4 to lock the bridge in the ground, by xing the deck to the abutments as in an integral
system that follows the ground motion with little additional deformation of its own.
2

Option 3 is not central in Part 2 of Eurocode 8, and is not dealt with in detail in this Designers
Guide. Option 4 is also a rather special case, applicable only to short bridges with up to three
spans but often with only one. It is dealt with separately in Sections 4.5, 5.4 and 6.11.3 of
this Guide. Options 1 and 2 are the main ones. The rst amounts to effectively isolating the
deck from ground shaking; it is treated in Eurocode 8 as seismic isolation. The second option
relies on ductility and energy dissipation in the piers. This chapter and most of the rest
focuses on these two options.
The features of conceptual design affecting the seismic behaviour and design of a bridge the most
are:
1
2

(for multi-span bridges) the continuity of the deck across spans over the piers
(for bridges with a concrete deck) whether the deck is monolithic with the piers.

To a certain extent these features are related to the method used for the erection of the deck (see
Table 4.1). It is very unlikely or even impossible for continuous decks to lose support and drop
from the piers. Even unseating from some bearings which is also unlikely will not have catastrophic consequences, and may be easily reversed. Unseating and dropping can be ruled out if the
deck is monolithically connected to the piers. In that case, the prevention of horizontal movement
between the deck and the top of the piers profoundly affects the seismic response of the bridge,
which is then dominated by the inelastic deformations and behaviour of the piers. It also affects
its seismic design, which is then based on the ductility of the piers. Monolithic or rigid connection
of the deck to the pier tops also affects the bridge performance under non-seismic actions. The
effect may be favourable (e.g. the performance under braking or centrifugal trafc actions in
railway bridges) or negative (e.g. the restraint of thermal or shrinkage deformations in a long
deck on stiff piers, which may even be prohibitive for the bridge).

4.2.

General rules for the conceptual design of earthquake-resistant


bridges

4.2.1

Deck continuity

Clauses 2.4(4), 2.4(9), The most important goal of the seismic design of a bridge is to keep the deck in place under the
2.4(10) [2]
strongest conceivable seismic action. In multi-span bridges, one of the risks to be faced by the
38

Chapter 4. Conceptual design of bridges for earthquake resistance

seismic design is local loss of support of the deck, due to unseating from a pier. The best way to
prevent drop of a part of a multi-span deck from one or more piers is by providing continuity of
the spans over all piers: a deck continuous from abutment to abutment. An exception may be
made in very long bridges of several hundred or over a thousand metres where intermediate
movement joints may be judiciously introduced. Such joints may be essential if it is considered
likely that a strong earthquake may induce signicantly different movement at the base of
adjacent piers, notably if the bridge straddles a potentially active tectonic fault or crosses
non-homogeneous soil formations. Intermediate movement joints may be placed within a
span as Gerber-type hinges with sufcient seat length. More often they are placed between
two spans whose ends are supported through separate bearings on the same pier. In such a
layout the movement joint should be wide enough to prevent pounding between the ends of
the two spans, in addition to providing sufcient support length against unseating (see also
Section 6.8.1.3).
Apart from breaking up the full continuity of the deck and increasing the chances of a drop-off,
intermediate movement joints increase the uncertainty of the seismic response: the parts of the
bridge separated by the movement joints (the frames of the bridge in US parlance) may
vibrate out of phase and experience pounding at the joints. Opening and closing of joints is a
nonlinear phenomenon, and capturing its effects may require a nonlinear analysis (normally in
the time domain). To take some of these effects into account without recourse to nonlinear
time-history analysis, the Caltrans Seismic Design Criteria (Caltrans, 2006) requires using, in
addition to a stand-alone model of each and every bridge frame between adjacent movement
joints, two global models that consider their interactions:
g
g

a compression model with all movement joints taken as closed


a tension model where movement joints are considered to be open and connected only
through the axial stiffness, EA/L, of any cable restrainers linking the deck in the
longitudinal direction across joint(s).

For bridges with several intermediate movement joints, Caltrans (2006) further requires the use
of several elastic multi-frame models, each one encompassing not more than ve frames plus a
boundary frame or abutment at each end; frames beyond the boundary ones are represented
by massless springs, and analysis results for boundary frames are ignored, while adjacent models
overlap by at least one frame beyond a boundary frame. The objective of this complex series of
analysis is to better capture the out-of-phase motion of frames and to account for the important
normal modes and periods of vibration of each frame without resorting to an unduly large
number of nodes from abutment to abutment. Despite its complexity, the above procedure
may not capture important features of the system response for the following reasons:
g
g

pounding between frames or the activation of cable restrainers linking them are unilateral
nonlinear phenomena that cannot be approximated by envelope linear models
if the bridge is long enough to have several intermediate deck separation joints, the effect
of the spatial distribution of the seismic ground motion may be quite important and worth
accounting for, even when such joints are provided.

As Part 2 of Eurocode 8 requires none of this analysis complexity, the sole reason for highlighting
here the provisions in Caltrans (2006) is to stress the uncertainty of the seismic response of
bridges with intermediate movement joints and the complexity of the analysis that this entails.
Reducing the uncertainty of the response is an important goal of conceptual design, and in
this case it is served well by avoiding such joints.
Deck spans composed of prefabricated girders, be they of concrete, steel or composite (steel
concrete), are normally simply supported on the piers. Adjacent spans can be connected by
encapsulating their ends in bulky cast-in-situ crossheads. However, this is not a common
practice. Normally, continuity of adjacent spans is pursued through a cast-in-situ topping slab
continuous over the joint between two girder ends (see also Section 5.5.1.4). Figure 4.1 depicts
an example. The slab should have sufcient out-of-plane exibility to allow different rotations
at the ends of the adjacent spans due to trafc, creep (and the ensuing moment redistribution)
and pier head rotation. This detail provides continuity of the pavement for motorist and

Clauses 2.3.2.2(4),
4.1.3(3) [2]

39

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 4.1. Precast girders simply supported on pier and connected for continuity via topping slab
Slab cast over the webs on a
2 cm expanding polystyrene layer

Cross beam

Bearings

Pier

passenger comfort and makes redundant intermediate roadway joints and the maintenance they
entail. It also ensures the continuity of seismic displacements of the deck and prevents impact
between adjacent spans under seismic actions. Finally, it serves as a sacricial seismic link
between the two spans against span drop-off after unseating. It is of note that the precast
girders of the twin 2.3 km-long Bolu viaduct were spared from dropping and triggering a
cascading collapse despite their unseating in the Duzce (TR) 1999 earthquake (Figure 4.2(a)),
due to their continuous topping slab. Part 2 of Eurocode 8 (CEN, 2005) makes specic reference
to such continuity slabs and their modelling.
It is normal practice to support prefabricated girders on a transversely stepped top surface of the
pier or the bent-cap or on corresponding concrete plinths, in order to achieve a transverse slope
Figure 4.2. Bolu viaduct in Duzce (TR) 1999 earthquake: (a) unseating of precast girders; (b) suspension
from the continuity top slab prevented span drop-off

40

Chapter 4. Conceptual design of bridges for earthquake resistance

of the top surface of the deck without differentiating the depth among the girders. By contrast, a
step of the pier top in the longitudinal direction is not a proper means to accommodate a difference in depth between two adjacent deck spans simply supported on the pier as, during the
longitudinal seismic response, the deeper of the two decks may ram the step. Instead, the
depth of the shallower deck should be increased over the support (through a deeper crosshead) to that of the deeper, so that both can be supported at the same horizontal level.
4.2.2
Uniform seismic demands on piers piers of different height
4.2.2.1 Introduction
For reasons of aesthetics, all piers or pier columns of a bridge usually have the same cross- Clauses 2.4(4), 2.4(6),
2.4(7) [2]
sectional dimensions. If the bridge has several piers with the same type of rigid connection to
the deck (i.e. all monolithically connected, or supporting it on xed bearings) as in option 2 of
Sections 2.3.1 and 4.1, differences in pier height are translated into differences in pier exibility
in a given horizontal direction (longitudinal or transverse), as exibility is approximately proportional to the third power of the pier height. This has certain implications for the longitudinal
or the transverse seismic response of the bridge. Some of these are unfavourable, and should be
avoided at the conceptual design stage, as highlighted in the following sections.
4.2.2.2 Conceptual design of bridges with piers of different heights for favourable
longitudinal response
The longitudinal inertia forces on an approximately straight deck (even one along which the
tangent to the axis does not change direction by more than 608) are about collinear. Owing to
the high axial rigidity of the deck, no matter where they originate, these forces are shared by
the individual piers (approximately) in proportion to their longitudinal stiffness. If the pier
columns have the same cross-sectional dimensions, shorter ones will undertake larger longitudinal seismic shears and develop higher seismic moments (which are approximately inversely proportional to the square of the pier height), requiring more vertical reinforcement than the rest.
This will further increase the effective stiffness of the shorter piers (cf. Section 5.8.1), and may
lead to a vicious cycle. In addition, regardless of the exact amount of their reinforcement, the
shorter piers will yield earlier and develop larger ductility demands than the others, possibly
failing sooner. Note that the shorter piers are normally towards the two ends of a long deck,
and, if rigidly connected to it, they constrain its thermal, creep and shrinkage deformations,
inducing in the deck high tensile forces and suffering themselves from the associated longitudinal
shears. The measures proposed below for the mitigation of non-uniform longitudinal seismic
demands in piers are quite effective in reducing these longitudinal constraints and their effects.
Conceptual design offers various ways around the problems posed by different pier heights:
g

If the height differences are rather small, the free height of the shorter piers may be
increased to be approximately the same as in all others. The added height may be in an
open (preferably lined) shaft under grade. The base of these piers should always be easily
accessible for inspection and repair of any damage, and above groundwater level.
Figure 4.3 shows an example.
If the pier heights are very different, rigid connection of the deck to the piers (monolithic
or through xed bearings) may be limited to a few piers of about the same height
normally the tallest ones. The deck may be supported on all other piers via bearings that
are exible in the longitudinal direction (elastomeric or sliding). Often, the tallest piers are
around the deck mid-length; so, this choice helps to relieve the stresses building up in the
deck and the piers due to the thermal and shrinkage movements of the deck in the
longitudinal direction. A typical example is the bridge in Figure 4.4. The deck is
continuous from abutment to abutment, with a total length of 848 m for the east-bound
carriageway and 638 m for the west-bound one, both with a radius of curvature of 450 m,
interior spans of about 55 m and end ones of about 44 m. It was cast span-by-span on a
mobile casting girder launched from pier to pier. Each deck is monolithically connected to
the ve centre-most and tallest piers, but tangentially sliding on the rest and at the
abutments (see Figure 4.5).
The cross-section of the shorter piers and of the upper part of the taller ones may be
chosen to present much smaller lateral stiffness in the longitudinal direction than the lower
part. In this way, the longitudinal stiffness of the piers can be balanced despite substantial
41

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 4.3. Construction of the lower part of the left pier of Votonosi bridge (GR) in a shaft for about
equal pier heights. (Courtesy of Stathopoulos et al. (2004))
L = 490.00

10

25.00

12

10

5.50

20.00

~47.00

5.50

~45.00

5.50

130.00
13.50

230.00
~13.50

130.00

10

20.00

10

25.00

12

differences in height. A usual choice for the longitudinally exible part of the pier is a
twin blade consisting of two parallel wall-like piers in the transverse direction. The lower
and stiffer part (possibly of very different heights in various piers) may be a hollow box.
Figure 4.6 shows a schematic, and Figure 4.7 a real example, of a balanced cantilever
construction (where the twin-blade piers offer additional advantages). Depending on the
relative length of its twin-blade and hollow box parts, plastic hinges may form either at
the very base of the pier or at the base of each of the individual columns of the upper
twin-blade part, or at both, one after the other. These possibilities should be taken into
account in the capacity design of the pier. All these potential plastic hinge regions
(including the top of the individual columns of the twin blades) should be detailed for
ductility.
If the deck is supported on all piers through elastomeric bearings, the stiffness of piers
with different heights may be harmonised by tailoring the total thickness t of the elastomer
so that the bearing stiffness Kb GA/t counterbalances the difference in pier stiffness, Kp ,
giving approximately the same composite stiffness from Eq. (D2.10) for all piers (see point
2 in Section 4.3.3.5). Note, however, that the large exibility of the elastomeric bearings
controls the horizontal stiffness of such bridges (see also Section 2.3.2.5 of this Guide).
If the section of pier columns is hollow, its thickness may be adjusted to balance the
difference in pier height and achieve either approximately uniform shear forces or
approximately uniform maximum moments among the piers. However, as the pier stiffness
is not very sensitive to the thickness of the hollow section, only small differences in the
pier height can be the accommodated in this way.

4.2.2.3 Transverse response of bridges with piers of different heights


The transverse inertial forces are distributed all along the deck. The seismic action effects they
induce in piers depend not only on their relative transverse stiffness but also on their tributary
deck length and the in-plane exural rigidity of the deck, which is normally high but does not
dominate the transverse response as much as the axial deck stiffness does for the longitudinal
response. Therefore, except in the special cases pointed out below, rigid transverse connection
of a long deck to all the piers produces a fairly uniform distribution of seismic demands
among them and is therefore acceptable, or even preferred. Exceptions to this rule are listed
below, alongside suggested conceptual design options.
42

Chapter 4. Conceptual design of bridges for earthquake resistance

Figure 4.4. Krystalopighi bridge (GR), with the deck monolithically connected to ve central piers and
free to move tangentially on ve or three piers near each end

Relatively short bridges (e.g. overpasses of three to ve spans) with transversely exible
piers and stiff abutments. If the intention is to resist the seismic action through ductile
behaviour of the piers (i.e. not as in an integral bridge), the deck should be transversely
unrestrained at the abutments.
Longer bridges, with very high transverse stiffness of the abutments and of the nearby
piers compared with the others. Rigid connection of the deck to all these stiff supports
may lead to a very unfavourable distribution of transverse shears among the supports, as
shown in Figure 4.9. The connection should be made transversely exible either at the
abutments or at the nearby piers.
Bridges with one or more secondary piers that serve the deck erection procedure.
Figure 4.10 shows an example of a balanced cantilever deck with a side span (on the left)
much longer than the central span. The side span is supported on an intermediate short
pier, sliding on it in both horizontal directions to avoid the unfavourable effects of a
transversely rigid connection.

4.3.

The choice of connection between the piers and the deck

4.3.1
Introduction: the effect of the construction technique
The fundamental choice is between connecting the deck monolithically with the piers and
supporting it on them through bearings xed (hinged, articulated) or movable (sliding or
elastomeric) or even via special isolation devices.
43

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 4.5. Piers for the bridge shown in Figure 4.4: (a) shorter ones near the ends of the deck,
supporting it without tangential restraint but with xity in the radial direction; (b) taller piers near the
centre, monolithically connected to the deck

3 4 = 12 piles
120
1
1

3 5 = 15 piles
120

For the type of bridges and the range of spans addressed herein, the piers are made of concrete.
They can be monolithic with the deck, but only if the deck is of concrete as well. If it is of steel or
composite (steelconcrete), it can only be supported on the piers via bearings.
The construction technique adopted for a concrete deck may dictate the choice between monolithic connection and support on bearings. As shown in Table 4.1, prefabricated girders are
normally supported on bearings (unless they are made integral with the top of the pier
through a cast-in-situ crosshead). Decks that are incrementally launched from the abutment(s)
are also supported on the pier tops during the launching through special temporary sliders appropriate for that operation. After the launching is completed, the deck is jacked up to replace these
sliders with the nal bearings. It is not practical to try there monolithic connections. Casting the
deck span by span on a mobile girder supported on and launched from the piers (as shown in the
upper right-hand corner of Figure 4.4) is convenient both for monolithic connection and for
support on any type of bearing. Concrete decks erected as balanced cantilevers are normally
monolithic with the pier, to stabilise the cantilever during construction. There are, however
Figure 4.6. Schematic of a bridge with a twin blade section in the upper 30 m of unequal piers and a
box section over the lower part (Bardakis, 2007)

1.50

44

7.40

7.00

7.40

7.00

10.00 30.00

30.00 30.00

144.00

1.50

144.00

91.00
10.00 30.00

144.00

30.00 30.00

91.00

Chapter 4. Conceptual design of bridges for earthquake resistance

Figure 4.7. Twin blade construction of the upper part of all piers of Arachthos bridge (GR) to reduce
and harmonise pier stiffness (see Figure 4.8 for the outer, shortest piers of the bridge)

cases, where a movable connection (with displacement and rotation in a longitudinal vertical
plane) is desired for the nal bridge, allowing, inter alia, seismic isolation. The deck may then
be supported on the pier during erection through bearings (usually temporary), with the deck
segment right above the pier head temporarily tied down to it (see Figure 4.8 for an example).
As another option, a span of a deck on bearings may be segmentally erected while cantilevering
out from a previously completed span, or with the segments suspended from a temporary pylon
through stays.
4.3.2
Monolithic connection
The simplest and most cost-effective connection of the piers to the deck is a monolithic one. From
the point of view of aesthetics, it is best to directly connect the pier column(s) to a cross-beam
incorporated within the depth of the deck, no matter whether it is a solid or voided slab, a
single- or multiple-box girder, or even a multiple T-beam deck. Monolithic connection of
a concrete deck to the piers is very common in Japan and the seismic prone areas of the
USA, but less common in Europe. The construction technique aside, it offers the following
advantages:
Figure 4.8. Temporary deck tie-down over the support on the shortest pier of Arahthos bridge (GR) via
four elastomeric bearings for stability during free cantilever erection (see Figure 4.7 for an overall
illustration of this bridge)

45

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 4.9. Abnormal distribution of seismic shears to supports (bottom) in bridge with transversely very
stiff outer piers and abutments (top)

Elevation
Plan

Wherever it can be applied, it is the cheapest and easiest connection of a concrete deck to
the piers, as the costs of special devices (bearings or isolation devices) and their inspection,
maintenance and possible replacement are avoided. Generally, it is preferable to an
articulated connection with bearings xed in both directions.
2 It is the best way to prevent unseating and drop-off or large residual displacements of the
deck that would be difcult to reverse in order to restore full operation of the bridge.
3 It lends itself best to design for ductile behaviour: energy dissipation and ductility can
develop not only at the bottom of the piers but at their tops as well (this does not apply to
single-column piers under transverse seismic action).
4 Fixity to the deck reduces slenderness and second-order effects in tall piers (except for
single-column piers in the transverse direction, if the deck is not laterally restrained at the
abutments).
1

There are serious disadvantages as well:


(a) For more than two spans, the piers restrain longitudinal movements of the deck due to
thermal actions and concrete shrinkage. As a result, signicant axial tension may build up
in the deck, alongside large bending moments and shears in the deck and in the piers
themselves. The longer the deck, the largest are the stresses due to the restraint of imposed
deformations. Ways to mitigate this problem have been discussed in Section 4.2.2.2.
(b) Monolithic connection precludes the use of seismic isolation and/or supplemental energy
dissipation.
(c) In the joint between the deck and the pier column, shear stresses due to the seismic
action are large; bond demands along pier vertical bars anchored there or along deck bars
passing through or terminating in the joint are high. Designing and detailing the joint for
these demands is not trivial.

Figure 4.10. Secondary short pier serving the erection of the long side span of Mesovouni bridge (GR)

46

Chapter 4. Conceptual design of bridges for earthquake resistance

4.3.3
Support on bearings
4.3.3.1 Types and structural functions of bearings
Supporting the deck on bearings is very common in Europe and Japan, but rather rare in the
seismic prone areas of the USA.
All types of bearings considered here have one common structural function: their articulation
capability. The vertical force to be transmitted is concentrated within the bearing support
area, within which the distribution of compressive stresses is essentially uniform. This is
effected by practically eliminating the rotational resistance within any vertical plane (articulation). To this end, the normal force is transmitted either through two concentric spherical
steel surfaces, one sliding over the other (in spherical bearings), or through a layer of resilient
material presenting a sufciently small rotational stiffness. This material can be either a laterally
encased non-laminated elastomer (as in pot-bearings) or a non-encased laminated elastomer
(in elastomeric bearings where the rotational stiffness is not negligible but acceptably low for
practical purposes).
Depending on their structural role, the following types of bearings are commonly used:
1

Fixed bearings, which are spherical or pot bearings that do not allow relative
displacement of the end bearing plates. Pot bearings are rather common in Europe, and
occasionally used in Japan, but rarely in the USA. Spherical bearings are more
uncommon.
Bearings sliding in all directions, to allow free displacement in any direction in the plane of
the bearing (plane of sliding). These are a fairly common type of movable bearing in
Europe or Japan, but rather rare in the USA. The bearing comprises one stainless steel
plate that can slide on another (the sliding plate), which is in turn anchored to the concrete
and has larger plan dimensions to accommodate the displacements. The sliding interface
material is usually lubricated PTFE, for low friction. Such bearings do not have rotational
capability. Sliding-cum-rotational capability can be achieved by adding a sliding plate to
one of the end plates of a spherical or pot bearing or to the end plate of an elastomeric
bearing pad (see below).
Bearings sliding in only one direction. This is effected by introducing a shear key guide
between the sliding plates of a bearing as above. Owing to the high contact pressure,
austenitic stainless steel and hard composite material are commonly used to form the
mating surfaces of the guide.
Laminated elastomeric bearings, presenting a low and elastic shear reaction to substantial
displacement in the plane of the lamination, according to their low shear stiffness,
Kb GA/tq (where A denotes the horizontal section area of the bearing, and tq and G the
total thickness and the shear modulus of the elastomer, respectively). These are the most
common type of bearing throughout the world (high damping rubber HDR is most
often used for them in Japan, to enhance energy dissipation). If the deck displacement is
governed by the low horizontal stiffness of elastomeric bearings (option 1 in Sections 2.3.1
or 4.1 of this Guide), the fundamental period of the bridge is lengthened and seismic force
demands are reduced. This is the simplest way of achieving seismic isolation, and is
recognised in Part 2 of Eurocode 8 as such.
Shear-key type of devices. Factory-produced devices termed shear keys or pins are fairly
common in Europe. They are capable of restraining the relative displacement in one or
both directions, transferring shear forces. The materials used for the mating surfaces are
the same as for the guides in bearings of type 3 above. As these devices do not transfer any
normal force, they are not bearings in the strict sense of the word. However, as they aim at
providing the other structural functions of xed bearings, they share their problems and
limitations highlighted in Section 4.3.3.4. They should allow free rotation about the two
horizontal axes and the vertical under various loadings or imposed deformations. Such
rotations, especially about the vertical axis, may be difcult to predict in long curved
bridges, erected and post-tensioned span by span, with different xed points at each
erection stage.

Least expensive of the above bearing types for the same axial load capacity are the elastomeric
bearings, with unanchored ones being much cheaper than anchored. Bearings sliding in any
47

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

direction come next, with those having an elastomeric component for freedom to rotate again
being less expensive. The cost of bearings xed in one or both directions increases with the magnitude of their shear force capacity. Spherical bearings are more expensive than pot bearings, but
present less resistance against and a higher capability of rotation.
4.3.3.2 Special seismic design requirements for bearings and limitations thereof
The problems and limitations highlighted here arise primarily, but not exclusively, from seismic
design requirements. The higher, therefore, the design seismic action, the more acute these
demands and limitations.
4.3.3.2.1 Seating on the deck
Sufcient overlap between supported and supporting elements should be provided at all movable
supports and directions of movement, unless xed bearings (or equivalent devices) are placed
there to preclude relative movement. The minimum overlap length per Part 2 of Eurocode 8 is
estimated according to Section 6.8.1 of this Designers Guide. Seismic links (e.g. via concrete
shear keys) are not required in addition. Fixed bearings or equivalent devices are designed to
carry the increased seismic action effects derived from capacity design (see case (b) in Section
6.3.2 of this Guide). Action effects derived directly from the seismic analysis can be used for
the design of xed bearings only if corresponding seismic links, capable of these capacity
design action effects, are additionally provided.
4.3.3.2.2 Transfer of horizontal forces to the concrete
The transfer of the vertical force of the bearing to the concrete follows the relevant rules of
Eurocode 2, and is normally taken into account in the generic design of the bearing by its
manufacturer. In the horizontal direction where relative displacement is not free, the design
shear force of xed bearings or equivalent devices and of anchored elastomeric bearings
should be safely transferred from the end bearing plate to the concrete. This is usually done
by means of anchor bolts, normally designed by the bearing manufacturer. The transfer to the
concrete of concentrated shear forces alongside the normal ones is not trivial, as reected by
the complexity of the relevant design rules in standards and guidelines (e.g. see CEN, 2008;
b, 2011). For bearings, this transfer is effected through the dowel and fastening action of
anchors, and depends heavily on the resistance along potential failure surfaces in the concrete
surrounding the anchor and extending to the free surface of the concrete element. The generic
bearing design cannot cover all the possibilities. This can only be done for each individual
design case.
The higher the design seismic shears to be transferred, the harder the design of the anchors.
High shear forces are quite usual, not only owing to capacity design but also because the use
of a single transversely xed bearing at the centre of each support is preferable, to avoid the
uncertainty about the distribution of transverse reactions to several xed bearings and the
restraint to imposed deck deformations. By contrast, usually two or more bearings are placed
for the transfer of vertical forces; for example, one bearing underneath each web of the deck
section, especially under the outer ones to minimise the vertical reactions due to the transfer
of torsional moments from the deck. So, in most cases the transfer of transverse horizontal
forces is separated from that of the vertical ones, using the special shear-key devices of point 5
of Section 4.3.3.1. As a result, the design seismic shears to be transferred to the concrete may
easily reach 50% or more of the corresponding vertical forces. If this transfer is not feasible,
vertical sliding bearings may be used instead to transfer the transverse horizontal forces. An
example is shown in Figure 4.5(a) for the tangentially movable supports of a bridge. This
solution is not problem-free either, as it requires provisions for the replacement of parts
subject to ageing, damage, contamination, etc.
4.3.3.2.3 Uplifting
Holding-down devices are normally needed to reliably prevent uplift of all the bearings at the
same deck support (on a pier or abutment) when there is no sufcient safety margin against
this risk in the seismic design situation. Individual bearings at one support do not have to
meet the relevant requirement of Part 2 of Eurocode 8 for increased safety. Nevertheless,
uplift under the action effects from the analysis in the seismic design situation should be
prevented at every single bearing.
48

Chapter 4. Conceptual design of bridges for earthquake resistance

4.3.3.2.4 Inspection, maintenance and replacement


All bearings should be accessible for inspection after an earthquake and under normal service
conditions, as well as for maintenance and partial or full replacement, if needed. To this
end, the design should provide for sufcient room around the bearing for the replacement
works, including the placement against the concrete surfaces of jacks, props and other temporary
means of support and strength. The bearing itself should also provide the appropriate separation
and connections between permanent and replaceable parts.
4.3.3.3 General advantages and disadvantages of supporting the deck on bearings
The construction technique aside, supporting the deck on bearings has its own advantages and
disadvantages. The pros and cons depend on the type of bearings used. On the pro side, no
matter the type of bearing:
Construction is in certain cases simpler than for monolithic connection.
Seismic stresses in the deck are diminished, compared with those in and around a
monolithic deck-column joint; for the longitudinal seismic action, they are almost
eliminated.
3 Single-column piers have similar behaviour in the longitudinal and transverse directions; if
spectral accelerations in these two directions are similar, pier seismic moments and shears
will be similar as well, allowing optimal pier design.
1
2

There are general cons as well:


(a) Bearings require inspection, maintenance and occasional replacement.
(b) If the pier comprises more than one column connected by a cap-beam (a bent cap in the
USA), its columns are not optimally used for earthquake resistance: in the longitudinal
direction they work as vertical cantilevers, while in the transverse direction their top is
nearly xed to the cap-beam.
Further pros and cons depend on the type of bearing, as highlighted below.
4.3.3.4 Fixed bearings
The advantages offered by xed (hinged, articulated) bearings are parallel to some of those of
monolithic connection:
To the extent they do not fail during the earthquake and are used over several piers, they
can prevent unseating and drop of the deck, as well as large residual displacements,
therefore facilitating restoration of full functionality of the bridge. However, they are a less
reliable means of achieving these goals than monolithic connection with some piers.
2 They are the only means, apart from monolithic connection with the deck, to mobilise the
ductility and energy dissipation capacity of piers. Of course, it is at the base (and not the
top) of only those piers that support the deck through at least one xed bearing that
ductility and energy dissipation develop.
1

They also share with monolithic connection cons (a) and (b) below:
(a) Signicant axial forces build up in the deck, alongside large bending moments and shears
in the piers, owing to the longitudinal restraint of deck thermal movements and concrete
shrinkage. In general, the problem is less acute than for monolithic deck/pier connection,
and can be mitigated as suggested in Section 4.2.2.2 in the second and third bullet points.
However, as the deck is free to rotate with respect to the pier in a vertical plane through
the longitudinal direction, the deck bending moments due to prestress and to a vertical
temperature difference component across the deck are lower.
(b) As with monolithic connection, it is not feasible to combine with seismic isolation and/or
supplemental energy dissipation.
In addition:
(c) Fixed bearings are in general more demanding for inspection and maintenance and more
costly than the other types of bearings, except for special isolation devices.
49

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Bearings that are xed in one horizontal direction (the transverse) and movable in the other (the
longitudinal) offer advantages 1 and 2 above in the direction where they are most crucial the
transverse while avoiding disadvantage (a) in the direction where it is relevant, namely the
longitudinal.
4.3.3.5 Elastomeric bearings
Laminated elastomeric bearings offer the following special advantages:
1

3
4
5

Owing to their very low horizontal stiffness, Kb GA/tq , they do not materially obstruct
the longitudinal movement of the deck due to thermal actions, concrete shrinkage or (for
prestressed decks) creep. However, near the ends of long continuous decks they may need
to be quite thick, in order to accommodate the deck longitudinal movements (see
disadvantage (c) below). To counter this drawback, sliding bearings (even elastomeric, with
a sliding end plate at the top) may be used near the ends of the deck and anchored
elastomeric ones around mid-length.
They offer a simple way to harmonise the stiffness of piers with different heights, by
tailoring the total thickness t of the elastomer so that the bearing stiffness Kb GA/tq
offsets the difference in pier stiffness, Kp , and give, in the end, about the same composite
stiffness as Eq. (D2.10) for the different piers.
Unless they fail (see disadvantage (f ) below), their behaviour is nearly elastic; they selfcentre after an earthquake, with little residual displacements.
They are of low cost and can be easily replaced.
They can serve as a simple means to achieve seismic isolation (and are recognised as such
in Part 2 of Eurocode 8): the fundamental period of the bridge is lengthened owing to the
low horizontal stiffness of the elastomeric bearings, and hence the seismic force demands
are reduced. In fact, they may be considered as the simplest and least expensive means for
seismic isolation. If the elastomer is not of HDR, the ensuing increase in horizontal seismic
displacement demands may be counterbalanced by supplemental (viscous) dampers. Such a
combination may be competitive with special isolation-cum-dissipation devices.
Unlike sliding bearings, whose friction properties strongly depend on the rate of loading,
the vertical stress and the condition of the sliding interface, the horizontal stiffness of
elastomeric bearings depends on well-dened and documented geometric and material
properties. Hence, the designer is fairly condent about the seismic action effects
transmitted by the bearing to the underlying pier or abutment as determined by the analysis.

Some of the disadvantages (notably those under (a) and (e) below) are the converse of the
advantages of monolithic connection or of xed bearings:
(a) They have such a high exibility compared with the pier that they accommodate almost
fully the seismic displacement demands, without letting the pier use any ductility and
energy dissipation capacity it may have. Hence, the behaviour is limited ductile (with
q 1.50) and the bridge is designed according to Chapter 7 of this Designers Guide for
bridges with seismic isolation. This, however, does not necessarily imply that the design is
not cost-effective.
(b) They are not a good choice over soft soils, because the ground motion may be rich in lowfrequency components due to site effects and/or the piers may tilt or displace differently
leading to deck unseating.
(c) To accommodate the deck longitudinal movements, they may need to be quite thick near
the ends of long continuous decks. As noted under advantage (1) above, this drawback
favours using them near deck mid-length and placing sliding bearings closer to the ends of
the deck.
(d) They are not appropriate or necessary when the piers are tall and the fundamental
period of the bridge would anyway be long even for rigid connection of the piers with the
deck.
(e) They cannot prevent the deck from falling (i.e. such supports are considered movable). To
avoid such a catastrophic event, Part 2 of Eurocode 8 requires the minimum overlap
length of Section 6.8.1 of this Designers Guide.
(f ) If they exceed their deformation capacity normally by debonding at the interface
between the elastomer and a steel plate, more rarely by toppling the deck will develop
50

Chapter 4. Conceptual design of bridges for earthquake resistance

large residual lateral displacements that will be hard to reverse in order to restore normal
operation of the bridge.
(g) They are subject to ageing; in addition, their behaviour during an earthquake may be
affected by past loading.
(h) They will tear under net vertical tension. To avoid this, Part 2 of Eurocode 8 includes
rules for holding-down devices.
4.3.3.6 Sliding bearings
Bearings with a horizontal sliding surface offer a single special advantage, in addition to the
general ones for bearings highlighted in Section 4.3.3.3:
1

As their friction coefcient is low under slow rates, they obstruct very little the
longitudinal movement of the deck due to thermal actions, concrete shrinkage or (for
prestressed decks) creep.

However, their drawbacks are several and serious. They share with elastomeric bearings the
limitations listed under points (a), (b), (d) and (e) in Section 4.3.3.5. In addition:
(a) They have no elasticity to restore large lateral displacements, therefore leading to
signicant and essentially unpredictable residual drifts, if not used alongside other elastic
devices capable of restoring displacements. As pointed out in Section 4.3.3.5 under
advantage (1) and disadvantage (c), the single pro of simple sliders may be combined with
pro (1) of elastomeric bearings, if the latter are used around mid-length of a long
continuous deck while sliding ones are used near its ends.
(b) They may be more expensive than elastomeric bearings.
(c) The value of their friction coefcient under dynamic rates is very uncertain and strongly
depends on the rate of loading, the vertical stress (increasing considerably when it is low),
the condition of the sliding interface, etc. If the friction coefcient is well above zero, the
bearing may induce signicant seismic shears in the pier below. If the value is very low,
the design of the bridge (seismic or not) cannot be condently based on the contribution
of a reliable lower bound of the force of the sliding bearings to the equilibrium of any
component of the structure (deck, pier or abutment). By the same token, such sliding
bearings are not reliable means of energy dissipation during the seismic response (see the
Clause 7.5.2.3.5(3) [2]
Note in Clause 7.5.2.3.5(3) of Part 2 of Eurocode 8).
(d) Apart from the large uncertainty in the value of its coefcient and its sensitivity to various
factors, friction is inherently a nonlinear behaviour mode: the direction of the force
changes with the sense of sliding although its magnitude may be constant. So, in theory, it
can be accounted for only through nonlinear analysis. However, this may not be possible
in practice, because a nonlinear analysis requires a well-dened set of external horizontal
actions (braking, wind or seismic) combined with imposed deformation actions (daily and
seasonal temperature variations, concrete shrinkage and creep). From the very large
number of possible scenarios for this sequence, it is evidently unfeasible to foresee the
worst one for the element considered. However, there are practical means to face these
difculties and achieve reliable results through simplied but safe-side estimation, based
on the following considerations:
As the friction forces of such sliding devices are relatively low, their direction should
always be selected so as to induce the worst effect for the case considered.
Although the magnitude of the friction coefcient mmax decreases with increasing
contact pressure sp (see Table 11 of CEN (2000)), the friction shear stress, mmaxsp keeps
increasing. Consequently, the maximum friction force corresponds always to the
maximum vertical force on the bearing.
So, both the magnitude and the direction of the friction force can be easily estimated in a
safe-side way for each case considered. In the seismic design situation in particular, the
following should also be borne in mind:
Clause 7.2.5.4(7) [2]
According to Part 2 of Eurocode 8 (Clause 7.2.5.4(7)), the value of mmax specied in
EN 1337-2 (CEN, 2000) may be used
The maximum normal and the corresponding maximum friction force of the bearing in
the seismic design situation are substantially lower than in the persistent-and-transient
51

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

one, because in the former the partial and combination factors applied on permanent
actions, Gk, and trafc actions, Qik , are equal to 1.0 or slightly above zero, respectively
(see Eq. (D6.1) in Section 6.2 of this Guide), while in the latter they are both markedly
greater than 1.0.
The seismic effects AEd may change the above picture. They comprise in general the
following components:
 The effect of the vertical component of the seismic action.
 The effect of the seismic action in the longitudinal direction (normally very small).
 The effect of the seismic action in the transverse direction due to the overturning
moment about the longitudinal axis. As far as the total friction force acting on a pier,
abutment or the deck is concerned, this effect is nil; by contrast, it may be quite
signicant (up to almost the full effect of gravity loads) for an individual bearing. If
the bearing is guided in the longitudinal direction, or it is a shear key device as
mentioned in point 5 in Section 4.3.3.1, the friction on the guide or the shear key due
to the transverse seismic shear reaction should also be added. For this case, CEN
(2000) species a value of mmax 0.2 for non-seismic design situations, no matter the
contact pressure. For the seismic design situation conditions, mmax should prudently
be increased to 0.3.
4.3.3.7 Special isolation bearings
It has been pointed out in Section 4.3.3.5 that simple elastomeric bearings are considered in Part
2 of Eurocode 8 as a means for seismic isolation. If their elastomer is HDR or they are used
alongside supplemental (uid viscous) dampers to counter the increase in displacements due
to the lengthening of the fundamental period, they may be considered to give a more
advanced isolation system like those addressed in this section. The latter combine exibility
(to lengthen the period and reduce the spectral accelerations and hence the design seismic
forces), with increased damping (to reduce the spectral displacements that accompany the
longer period). The damping may be provided by the isolation bearing itself (as in HDR
bearings, in leadrubber bearings, in sliders with a horizontal or spherical sliding surface and
higher and well-controlled friction properties, etc.), or by supplemental devices that intervene
between the deck and the top of the pier (or abutment) and are subjected to their relative
displacements without supporting vertically the deck. They may be uid viscous dampers,
ductile steel units dissipating energy through plastic deformations, special magneto-rheological
dampers, etc.
Although the specic advantages and disadvantages depend on the isolation system chosen, some
general observations may be made. On the advantage side:
Like simple elastomeric bearings or sliders, there is very little obstruction to longitudinal
movement of the deck due to thermal actions, concrete shrinkage or (for prestressed decks)
creep.
2 If the necessary skills are available and the necessary design effort is made, seismic
isolation may be the most cost-effective seismic design option, especially in high-seismicity
regions.
3 The devices themselves are subject to much stricter quality control than simple bearings,
and their properties are better known and documented.
4 Isolation systems (at least if designed to Part 2 of Eurocode 8) have self-restoring features
that allow the deck to sufciently recentre even after very strong shocks.
1

Special isolation systems share with elastomeric or sliding bearings the limitations listed under
(a) and (b) in Section 4.3.3.3 and (b), (d), (e) in Section 4.3.3.5. A list of additional ones follows:
(a) They require special expertise and experience in design and analysis (which is normally of
the nonlinear response-history type) and suitable analysis software.
(b) The cost of the devices themselves is relatively high; however, with the necessary design
effort and a judicious choice of the type of isolation system and its devices, the total cost
of the bridge may well be less.
(c) There is some uncertainty regarding the behaviour of the devices and of the bridge as a
whole in case the displacement capacity of the device is exceeded. To counter this, Part 2
52

Chapter 4. Conceptual design of bridges for earthquake resistance

of the Eurocode 8 requires increased reliability for the displacement capacity of the
isolation devices, by multiplying the seismic displacement demand from the analysis by a
factor of 1.50.
4.3.3.8 Concluding remarks
As repeatedly pointed out and exemplied in Section 4.3.3, the designer may choose to have one
type of connection or bearing at certain locations under the deck and to combine it with one or
more other types elsewhere, in order to maximise the benet from their advantages and counter
their drawbacks to the maximum possible extent. One example is the use of monolithic connection or elastomeric bearings around mid-length of long continuous decks and of sliding ones near
the ends. As also noted (certain), bearings may be chosen to be xed in one horizontal direction
and movable in the other. In choosing among the various options available, the designer should
keep in mind that the connection of the deck as a whole to the piers may be considered as a
parallel system, where the stiffer and stronger component(s) control the behaviour and the
displacements of the bridge.

4.4.

The piers

4.4.1
Sections of pier columns for efcient detailing
4.4.1.1 Solid circular columns
Solid circular pier columns are quite common around the world. A circular section has the same
strength and rigidity in every horizontal direction. So, it seems ideal for pier columns that work
as vertical cantilevers in both directions (e.g. when the deck is supported on the pier through xed
or horizontally exible bearings), or belong to multi-column piers monolithically connected to the
deck. In addition, it lends itself better than any other section to the efcient connement of the
concrete and antibuckling restraint of vertical bars in this case through circular hoops or a continuous spiral, with little need for cross-ties or link legs at right angles to the perimeter of the section.
A circular perimeter hoop or spiral contributes to the shear resistance of the section in diagonal
tension with only p/2 1.57 tie legs, not two. However, this is not a serious drawback of circular
pier columns, because they are commonly quite slender and hence not critical in shear. Also, if
shear resistance in diagonal tension is critical, interior rectangular links or straight cross-ties
may be added, engaging vertical bars across the full section. Such tie legs contribute to the
shear resistance in diagonal tension with their full cross-sectional area, to be contrasted with
circular ties contributing with a fraction of p/4 0.785. This is less material consuming than a
reduction in the spacing of hoops or spirals or the arrangement of the vertical bars in two
layers with a second hoop in addition to the perimeter one in-between.
Solid circular sections with a large diameter (e.g. over 34 m) are normally avoided, not only as
they are not cost-effective compared with hollow circular ones but also because of concerns about
their large volume of unreinforced concrete.
It is sometimes considered harder to mesh the two-way top reinforcement of a pile cap or spread
footing or the bottom one of the deck or a column cap-beam with the dense vertical reinforcement of a circular column, compared with that of a column with sides parallel to those of the
reinforcing mesh in the foundation element.
4.4.1.2 Solid rectangular columns
Compact solid rectangular pier columns are common in Japan, but less so in Europe, and rare in
the USA. The detailing of large sections for plastic hinging is not cost-effective compared with
circular sections, because it requires many long cross-ties or link legs at right angles to the
sides in order to mobilise the straight perimeter stirrup for connement of the concrete and to
restrain against buckling the large number of vertical bars arranged along the side. This is
feasible only if the section is relatively small and does not require unduly long cross-ties or intermediate link legs. If a large square section is chosen over a circular one (e.g. for convenience of
the formwork), its vertical bars may be arranged in a ring and engaged by circular hoops or a
circular spiral. The four corners of the section are unreinforced and would rather be chamfered.
This idea may be extended to solid rectangular sections with a ratio of the two sides of around
1.5:1, employing interlocking circular hoops or spirals. Again, the four unreinforced corners
of the section are chamfered.

Clauses 6.2.1.2,
6.2.1.4(4) [2]

53

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

4.4.1.3 Wall-like piers


Rectangular wall-like piers, with the long dimension in the transverse direction of the bridge, can
provide continuous support almost all along the width of the deck. So, they are quite convenient
under thin concrete decks. They are popular in Japan and becoming increasingly so in Europe,
but not so much in the USA. Owing to the large cross-sectional area, the level of vertical stress is
quite low, and hence connement is not essential, especially in the weak direction of the pier,
where the shear-span-to-depth ratio is high and favourable for ductility. Nevertheless, it can
be conveniently provided by engaging pairs of vertical bars along the two long sides of the
section by fairly short cross-ties or hoop legs at right angles to the long side of the section.
The intermediate vertical bars along the two short sides are normally few; they can be laterally
restrained through short oblique cross-ties or link legs also engaging nearby bars on the long
sides.
In the longitudinal direction of the bridge, such a wall is quite ductile owing to its slenderness.
This may not be the case in the transverse direction. Indeed, if in that direction the shear span
(moment-to-shear ratio) is less than three times the depth of the section, and the pier top is
rigidly connected to the deck, the q factor for ductile behaviour is decreased to reect the
reduced ductility (cf. Table 5.1 in Section 5.4 of this Guide). However, the moment and shear
resistances in the strong direction of such a pier are large enough to resist the design seismic
action effects even for a reduced q factor.
The main drawback of wall-like piers is the very high design seismic action effects they may
deliver to the foundation. If derived from capacity design, these action effects reect the large
overstrength of the plastic hinge in the pier, leading to almost elastic seismic action effects in
the foundation (i.e. with q 1.5). An option that may prove more cost-effective is to use piles
for the foundation and allow plastic hinges in them for seismic design in the strong direction
of the pier (using the value q 2.1 in this case).
Clause 6.2.4(2) [2]

Clauses 6.2.4(2),
6.2.4(3) [2]

Pairs of parallel pier columns with a wall-like section (twin blades) are often used to support
decks built as balanced cantilevers (see Figures 4.6 and 4.7). The pair of pier columns is quite
effective in stabilising the deck during free cantilever construction, but owing to the slenderness
of the individual columns have low longitudinal stiffness and reduce the longitudinal stresses
that develop in the deck due to restraint of thermal or shrinkage strains. As pointed out in the
last bullet point of Section 4.2.2, if the various piers of the bridge have very different heights
(as in Figures 4.6 and 4.7), all the piers may be chosen to have this type of twin-blade section
over a certain length of their upper part, while the lower part of very different height among
the piers may have a hollow box section (see Figure 4.6).
4.4.1.4 Hollow rectangular piers
Tall piers have by necessity a large section. Large sections are normally constructed as hollow, to
increase the strength and the rigidity (against second-order effects as well) for a given volume of
concrete. The reduction in pier weight and mass for a given strength and rigidity decreases
inertia forces and the vertical loads on the foundation. Hollow rectangular sections are quite
common, especially in Europe and Japan.
The vertical bars are arranged all along the outer and inner perimeters, with particular concentration at the corners. Pairs of vertical bars across the thickness of the box section are engaged by
fairly short cross-ties or at the corners of closed links placed within the thickness of the section
(see Figure 4.11 for examples). Part 2 of Eurocode 8 requires a web thickness in the plastic hinge
region of at least one-eighth of the clear distance of the webs framing in the other direction, to
prevent local buckling when the web acts as compression ange. When the pier is slender in
one horizontal direction, the thickness of the pair of webs parallel to that direction may be
controlled by the provisions of Eurocode 2 for slenderness and second-order effects. If it is
not slender, ultimate limit state (ULS) resistance in shear (against diagonal compression) may
govern the web thickness. Note also that, at least for the most common layouts of transverse
reinforcement shown in Figure 4.11, only four link legs run the full depth of the section in
each horizontal direction, and count towards the shear resistance of the section in diagonal
tension.

54

Chapter 4. Conceptual design of bridges for earthquake resistance

Figure 4.11. Examples of reinforcement arrangement in box piers. (Courtesy of Mechaniki SA)

It is possible to vary the thickness of the section along the pier or between piers to suit better the
strength and stiffness needs of the pier and the bridge, without changing the outer dimensions
and appearance of the pier.
What has been said for hollow rectangular piers can be readily extended to polygonal hollow
piers.
4.4.1.5 Hollow circular piers
Hollow circular (annular) sections are less common for tall piers than hollow rectangular or polygonal ones (especially in Europe). The main problem for the use of hollow circular sections in
ductile piers is that circular hoops cannot conne the inner face of the section: the circumferential
tensile force in such hoops produces radial deviation stresses pointing away from the conned
concrete core, and may cause loss of the cover concrete and implosion of the inner face of the
section. So, the hoops around the inner face can serve only as shear reinforcement. To conne
the inner face, radial cross-ties or link legs should be placed across the thickness of the
annular section, as in a hollow rectangular one. Alternatively, the thickness should be large
enough to keep the strain at the inner face below the ultimate strain of unconned concrete
(1cu 0.35%) even when the ultimate curvature is attained and after spalling of the concrete
cover at the outer face. The lower limit of one-eighth of the inner diameter imposed by Part 2
of Eurocode 8 on the thickness of ductile annular piers against the occurrence of local wall
buckling unintentionally helps in protecting the inner perimeter of the section from large
strains at ultimate curvature. Note, though, that radial cross-ties or link legs across the thickness
are useful to supplement the outer face hoops, if the large diameter of the section reduces their
effectiveness for connement and against buckling of the vertical bars.

Clause 6.2.4(3) [2]

As in hollow rectangular piers, the thickness of the section may vary up the pier or between piers
to adapt the stiffness and the resistance of the pier to the seismic demands without affecting the
aesthetics of the bridge.
4.4.2
Single- versus multi-column piers
Except for the twin-blade piers often used in balanced cantilever bridges for the reasons
discussed in previous sections, tall piers are normally made of a single column, to increase
rigidity and strength and reduce slenderness for given concrete volume. When the pier is relatively short and the deck wide or shallow, the designer has the option of using either a walllike pier (with the limitations pointed out in Section 4.4.1.3) or a pier with more than one
column across the deck.
If bearings of any type xed, hinged, sliding, elastomeric or even isolation bearings support
the deck at the top of the pier, a single-column pier is an efcient choice, as it is effectively a
55

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

vertical cantilever in any horizontal direction. If there is more than one bearing across the deck, a
capital or a hammerhead on top of a single-column pier can accommodate them. If the bearings
present the same stiffness in the transverse and the longitudinal directions, a single circular pier
will be subjected to very similar seismic shears and moments in these two directions, and may be
very cost-effective.
The single-column option is simple and the behaviour is clear, but the redundancy is low.
If the pier is monolithically connected to the deck, then in the longitudinal direction all of its
columns have their top nearly xed and a shear span slightly longer than half their free height
(see Eq. (D5.4b) in Section 5.4 of this Guide). If the pier has a single column, in the transverse
direction it works as a vertical cantilever in fact, its shear span is longer than its free height
owing to the rotational inertia of the deck (see Eq. (D5.5) in Section 5.4). Such a singlecolumn pier should have a signicantly larger resistance and rigidity in the transverse than in
the longitudinal direction, and is not a very cost-effective option. If two or more columns are
used across the deck, they will all have the top essentially xed against rotation. Their shear
spans will be similar in the transverse and longitudinal directions, and their design seismic
moments for transverse seismic action will be minimal (equal to the seismic shear force multiplied
by one-half the column clear height, in lieu of the full clear height in a single-column pier).
Additionally, energy dissipation and ductility demands will be shared by plastic hinges at the
top and the bottom of each column, for both transverse and longitudinal seismic action.
Needless to say, a circular cross-section is the optimal choice for multi-column piers.
As pointed out in Section 4.3.2, if the pier (column) is monolithically connected with the deck,
aesthetics suggest a connection with a cross-beam that is fully incorporated within the depth
of the deck. If the deck is shallow without deeper and protruding cross-beams, it is not easy to
verify and detail its joints with pier columns for ductile behaviour of the bridge with plastic
hinges forming at column tops, unless the columns are also small and their number larger.
This will reduce the proportion between the moment transferred by each connection and the
effective volume of the connection (which extends into the deck beyond the perimeter of the
column).
Connecting the head of a multicolumn pier to the deck through xed bearings is not as efcient as
a monolithic connection. Elastomeric bearings are preferred. A continuous slab or box-girder
deck may be directly supported on the column tops through elastomeric or seismic isolation
bearings. If the deck consists of precast girders, the bearings are usually supported by a transverse
cap-beam framing at the tops of the columns (forming a bent in US terminology), with frame
action in the transverse direction and free-cantilever action in the longitudinal. The cap-beam is
normally deep, to allow straight anchorage of the column bars within its depth, and hence stiff.
So, the top of closely spaced columns is nearly xed against rotation in the plane of the multicolumn pier. When the cap-beam supports only horizontally movable bearings, the multicolumn pier should be designed for limited ductile behaviour with q 1.5. Should even one of
the bearings be xed, the multi-column pier should be designed and detailed for ductile behaviour
with a q value larger than 1.5. At any rate, a multi-column pier supporting the deck through a line
of bearings may be a simple choice, but is not so cost-effective. A layout that should be avoided
for the piers is a portal frame of two columns spaced apart by more than the deck width,
supporting a cap-beam that is integral with the deck (i.e. fully or partly within the depth of
the deck). Although such a layout is common in the USA and occasionally used in Japan,
Part 2 of Eurocode 8 does not address it. A major source of uncertainty and a weak link in
such a portal frame is the outriggering length of the cap-beam between the deck and the pier
columns.
Overpasses with three to four spans are most often supported on piers with several slender
columns each, which may be hard to design for resistance against impact of heavy trafc vehicles.
A multi-column pier (be it monolithic with the deck or supporting it through a string of bearings
on a cap-beam) at a skew to the bridge provides a more complex support condition to the
deck than a single-column pier. A skew bridge may be avoided over a skewed crossing, if the
abutments are placed at right angles to the deck axis and single-column piers are used.
56

Chapter 4. Conceptual design of bridges for earthquake resistance

4.4.3
Sizing of pier columns
4.4.3.1 General sizing criteria
To complete the conceptual design phase, once a section shape is chosen for the pier columns,
their size should be selected. This section covers general criteria for the selection of the
column size, pertaining to general design situations, including seismic ones for bridges of
limited ductile behaviour. Section 4.4.3.2, by contrast, refers specically to limitations imposed
by seismic design for ductile behaviour.
As repeatedly said in this chapter, for aesthetic reasons normally all the pier columns of a bridge
have the same outer dimensions. Choosing the same shape for different piers or pier columns but
different cross-section sizes may be aesthetically worse than having a different shape altogether.
General criteria for the determination of the dimensions of the pier columns include:
g
g

practical considerations, taking into account the construction technique and procedure for
the piers themselves and the deck, as well as the connection to the deck
the slenderness of the column, in order to limit second-order effects under factored gravity
loads (in the persistent and transient design situation of the completed bridge), or,
preferably, to be able to ignore them as small.

Concerning the construction procedure, often intermediate stages before the deck completion
may be more critical for the stability and design of the piers. In other cases, the geometry of
the piers is conditioned by the need to accommodate and support special and heavy equipment
for the erection of the deck. That said, for the bridge to be cost-effective, the choice of its structural system should strike a rational balance between the requirements of design situations
pertaining to intermediate construction stages and to those of the completed bridge; the latter
should normally govern the design.
A column section depth greater than that of the pile cap or spread footing underneath may make
it harder to verify their joint region for the capacity-design action effects of Section 6.4.4 of this
Guide. By the same token, a column monolithically connected to the deck should have a section
depth in the plane of xity to the underdeck (i.e. in the longitudinal direction, and for multicolumn piers in the transverse as well) less than the deck itself. If the deck is supported on
bearings, the pier top should be sufciently wide to accommodate the bearings, the required
seat lengths (see Section 6.8), any shear keys and the clearance to them. To this end, a hammerhead or a capital may be added to the pier top, or the pier may be ared upwards.
Slenderness considerations are crucial for tall piers. In a bridge designed for earthquake
resistance it is natural to size the pier columns so that second-order effects in the persistent
and transient design situation may be ignored altogether as small. Clause 5.8.3.1(1) in
EN 1992-1-1:2004 (Part 1-1 of Eurocode 2) gives the upper limit for the column slenderness
above which second-order effects should be taken into account in the fairly detailed and
complex way prescribed in Eurocode 2. The effective length of the pier column (i.e. the numerator
in the column slenderness) is a prime factor in these sizing calculations. In the usual case of a deck
simply supported on the abutments and free to translate there in the longitudinal direction, the
effective length of pier columns may be approximated as follows:
1

For monolithic connection with the deck:


(a) in the longitudinal direction: the full clear height of the pier column (i.e. column top
free to translate horizontally without rotation)
(b) in the transverse direction:
if the deck is laterally restrained at the abutments and has high in-plane rigidity:
one-half the clear height of the column (i.e. top xed against rotation and transverse
translation)
if the deck is free to translate laterally (i.e. it is unrestrained at the abutments, or is
very long or has low in-plane rigidity):
 for single-column piers: twice the clear height of the pier (i.e. free cantilever)
 for multi-column piers: the full clear height of the pier column (i.e. column top
free to translate horizontally without rotation).
57

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

For a deck supported on bearings (xed or movable):


(a) in the longitudinal direction: twice the clear height of the pier column (i.e. free
cantilever)
(b) in the transverse direction:
if the deck is laterally restrained at the abutments and has high in-plane rigidity:
 for columns directly supporting the deck: 70% of the clear height of the pier (i.e.
top free to rotate but restrained against horizontal translation)
 for multi-column piers framing into a stiff cap-beam: one-half of the clear
height of the column (i.e. column top xed against rotation and transverse
translation)
if the deck is free to translate laterally (i.e. it is unrestrained at the abutments, or is
very long or its in-plane rigidity is low):
 for single-column piers: twice the full clear height of the pier (i.e. free
cantilever)
 for multi-column piers framing into a stiff cap-beam: the clear height of the pier
column (i.e. column top free to translate horizontally, without rotation).

If the deck is integral with at least one abutment (see Section 4.5.3), the rst bullet point under
cases 1 (b) and 2 (b) applies not only to the transverse but also to the longitudinal direction of the
bridge.
The above refers to the column effective length in the completed bridge. In a free cantilever
bridge, the effective length of single-column piers during deck erection is twice the clear height
of the pier (i.e. a free vertical cantilever) both in the longitudinal and in the transverse directions;
if the upper part of the pier consists of two parallel wall-like columns (twin blades, see
Figures 4.6 and 4.7), the clear height of the column applies as its effective length in the longitudinal direction. Critical for pier slenderness may then be the stage after completion of the hammer
and before connection at mid-span or at an abutment with other completed parts of the bridge.
The permanent loads applied at that time and the value of the creep coefcient at that age of the
pier concrete and not at the end of the life of the bridge should be used in the upper limit for
column slenderness above which second-order effects should be taken into account as per clause
5.8.3.1(1) of Part 1-1 of Eurocode 2; this differences may offset the longer effective length of the
pier at that stage.
A lower limit to the cross-section of non-slender pier columns monolithically connected to the
deck may be imposed by the verication of concrete for the fatigue ULS according to
clause 6.8.7(101) of Part 2 of Eurocode 2 (equivalent damage stress) for railway bridges,
or clause 6.8.7(2) of Eurocode 2-Part 1-1 (simplied verication based on the frequent
combination) for roadway bridges.
4.4.3.2 Specic criteria for bridges with ductile behaviour
The choice between ductile or limited ductile behaviour (including seismic isolation) for the
bridge has a decisive impact on the sizing of its piers and vice versa, as plastic hinges for
energy dissipation zones are allowed only in the piers. If design for ductile behaviour is
chosen, only the pier plastic hinges are dimensioned to resist at the ULS the bending moment
and axial force from the analysis for the seismic design situation. All other sections of the pier
and of all other components of the bridge structure are dimensioned to resist at the ULS the
capacity design effects, computed according to Section 6.4 of this Guide from the design
moment resistance of the pier plastic hinges. Should that moment resistance substantially
exceed what is required according to the analysis results for the seismic design situation, all
other parts of the bridge (including the pier itself in shear) are penalised. The moment resistance
of a plastic hinge section may be unduly large if:
g
g

the dimensions of the section are unnecessarily large


the vertical reinforcement ratio is governed by non-seismic design situations (including at
intermediate stages of the construction) or by detailing and minimum reinforcement
requirements.

Therefore, for cost-effective seismic design for ductile behaviour:


58

Chapter 4. Conceptual design of bridges for earthquake resistance

The cross-section of the piers should not be larger than rationally needed, and the vertical
reinforcement of their plastic hinges should not be beyond what allows reliable
construction.
The nal vertical reinforcement of the pier plastic hinges should be as close as feasible to
what is required for the biaxial bending moments and the axial force from the analysis for
the seismic design situation.
The q factors to be used in the longitudinal and transverse directions should be as close as
feasible to the maximum values allowed in Part 2 of Eurocode 8, listed in Table 5.1 of
Section 5.4. To this end:
the location of the expected plastic hinges should be accessible
the shear span ratio of the pier, Ls/h, should not be less than the value of 3.0, below
which the q factor is penalised (see Section 5.4 of this Guide)
the pier maximum axial load ratio, hk NEd/Ac fck , in the seismic design situation
should not exceed the value of 0.3, above which the q factor is reduced (see Section 5.4).

The axial load ratio, hk , has a number of important inuences on seismic design, especially for
ductile behaviour, because it is important for the ductility of the pier column. If it is high, the
connement requirements in the plastic hinge are larger, especially in piers designed for ductile
behaviour (see rows 18, 19 and 22, 23 in Table 6.1 of Chapter 6). In addition, the overstrength
factor applied to the moment resistance of ductile piers for capacity design calculations increases
with increasing hk (see Eq. (D6.6b) in Section 6.4.1 of this Guide). More importantly, as already
pointed out, the q factor of bridges designed for ductile behaviour is controlled by the maximum
value of the axial load ratio in any pier column of the bridge (see Section 5.4). In all of these calculations the value of hk is that determined from the maximum axial load in the pier column,
NEd , in the seismic design situation. In multi-column piers and for the transverse seismic
action, the overturning moment may signicantly increase NEd over the value due to the
gravity loads alone. That said, very seldom does the level of hk control the area of the pier
section: the size of pier columns selected on the basis of practical or slenderness considerations
normally gives quite low values of hk in the seismic design situation.

4.5.

The abutments and their connection with the deck

4.5.1
The role of abutments
The abutments have a dual role:
1
2

to provide vertical support to the deck, like the piers, but at its very end
to act as a retaining wall for the backll beyond the end of the bridge.

The second role normally governs the form, dimensions and cost of the abutment. In fact, an
abutment has a substantially higher cost than a typical pier, despite its much lower vertical
reaction from the deck and shorter height. Cost is governed by the high, nearly permanent,
earth pressures of the backll, the resultant of which and the impact of the retaining role are proportional to the square of the height of the retained ll. The main consequences of the retaining
role are:
g

The horizontal stiffness of the abutment is much larger than that of a pier:
in the longitudinal direction owing to the contribution of the backll and/or the
required resistance to earth pressures
in the transverse owing to the wall-like shape of the abutment.
It is benecial and cost-effective to brace longitudinally the abutment at the level of the
deck against the action of earth pressures, especially if the depth of the backll is large. A
monolithic connection of both abutments to the deck is very effective to this end, as it
mobilises in the bracing action the opposite abutment and the reaction of the backll
behind it.

To serve its second role, an abutment is usually a deep retaining wall on a footing on competent
ground or on the cap of piles going down to competent and non-liqueable soil. Sometimes the
retaining wall consists of a single curtain of dense piles supporting a shallow beam on which the
deck is seated. A deep continuous retaining wall or a curtain of dense piles may be pushed
inwards by a laterally spreading or liqueed backll. If the nature of the backll does not rule
59

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

out this possibility (e.g. at a river or lake bank), the piles supporting the shallow seat beam would
be better sparse and deep, for uninhibited lateral ow of the ground between them.
4.5.2
The connection options
As far as the seismic response and design are concerned, the abutments may play a signicant or a
minor role, depending on how they are connected to the deck in the horizontal direction. The
main connection options are:
1
2

to have a deck integral with the abutments


to seat the deck on the abutments through bearings movable only in the longitudinal
direction or in both horizontal directions.

Option 1 can be adopted only if both the deck and the abutments are of concrete. They then move
in unison in any horizontal direction. For option 2, the designer may choose:
2a to let the deck free to move horizontally at the abutment
2b to restrain horizontally the deck at the abutment
2c to let the deck move until it comes in hard contact with the abutment and then move with

it (or not move at all).


Option 2b is not very different from option 1. Unlike the integral connection, which refers to both
the longitudinal and the transverse directions and has similar implications for the design of the
bridge in both, one option among 2a to 2c may be adopted in the transverse direction and
another in the longitudinal. More details are given in the sequel.

Clause 6.7.3(1) [2]

4.5.3
Deck integral with the abutments
The connection of the abutment to the deck is considered as rigid if it is monolithic (integral), or
if in both horizontal directions (longitudinal and transverse) it is effected via xed bearings or
links designed to carry the seismic action. Unlike in bridges with a movable connection of the
deck to the abutments, the abutments of integral bridges play a major role for the seismic
resistance in any horizontal direction.
As pointed out in Section 4.5.1, the top of abutments integral with the deck is braced horizontally
in a very reliable way, with considerable benets against seismic actions and earth pressures from
the backll. The absence of movement joints improves motorist comfort especially over bridges
of short length and reduces initial and maintenance costs through savings on bearings and
roadway joints. These cost savings are a prime advantage of integral bridges. In railway
bridges, transverse movement of the deck with respect to the abutment is avoided and tracks
are protected.
Integral connection is the best way to prevent the deck from dropping from the abutment. This is
especially important if the deck is very skewed to the abutments or sits on non-parallel ones (i.e. if
the geometry is kinematically favourable for unseating of simply supported decks see the last
two paragraphs of Section 6.8.1.4 of this Guide). Note that, if the end supports are very skewed,
integral connection offers the advantage of eliminating the high and uncertain concentration
of vertical reactions to gravity loads that takes place under the obtuse angle of bridge decks on
bearings. Last, but not least, if headroom below the deck is at a premium, a deck working with
the abutments as a portal frame may be made much thinner than a simply supported one, owing
to its reduced span moments.
The above advantages may be offset by a serious drawback: integral abutments restrain longitudinal deformation of the deck due to thermal actions and concrete shrinkage. As a result,
large tensile forces may build up in the deck, alongside longitudinal forces in the abutments
and their foundation. Additionally, the deck cannot be longitudinally post-tensioned, without
losing the (large) part of the prestressing force that goes not to the deck itself but to the
abutments (and from there to the ground, with undesirable results): the deck may prot only
from the load-balancing effect of the deviation forces of curved tendons, but very little from a
longitudinal compressive force. Note also that the high bending stiffness of the system
abutment-cum-backll prevents efcient use even of these deviation forces. For these reasons,

60

Chapter 4. Conceptual design of bridges for earthquake resistance

only short decks are built integral with the abutments. They mostly have one span, or sometimes
two or three short continuous ones, but not more than four. Their total length is normally
much less than 100 m. Also, normally such a deck is not post-tensioned. For such bridges,
accounting for a large fraction of the bridgework along motorways (at almost every short
under- or overpass), an integral connection is the solution of choice, and a very robust one
indeed.
As the seismic behaviour and design of integral bridges is normally dominated by the abutments,
any intermediate piers have almost no contribution to the longitudinal earthquake resistance.
Transversely, they have a certain contribution if they have a wall-like section and the deck is
relatively narrow and long (a rare combination for this type of bridge). A row of slender
(e.g. circular) columns, monolithically connected to the deck, may then be optimal for the
intermediate piers.
It makes no sense to support the deck on both abutments through xed bearings (option 2b in
Section 4.5.2) instead of building it as integral: the main drawback of integral connection
remains, while the span moments are not reduced, and deck drop-off in case the xed bearings
fail cannot be precluded. Additionally, the xed bearings should be designed for high horizontal
forces due to the restraint. There is some scope, though, for a deck integral with one abutment
and supported on the other through horizontally exible bearings. This option relieves the
bridge from stresses due to restraint of imposed deformations, allows longitudinal prestressing
and decreases signicantly the chances of deck drop-off compared with a deck simply supported
at both ends. However, the seismic response and behaviour is more uncertain and the analysis
more complex.
Clauses 4.1.6(9),
As explained in more detail in Section 5.4 of this Guide, bridges often having the deck integral
with the abutments may be considered to be locked-in the ground, and follow its horizontal 4.1.6(10), 6.7.3(1)(4),
6.7.3(9) [2]
motion without amplifying it much. If their fundamental period T is less than 0.03 s, Part 2
of Eurocode 8 allows for them to be designed as elastic (i.e. with q 1) but rigid (i.e. with
forces equal to the masses multiplied by the design peak ground acceleration). If, however, T
is longer than 0.03 s, Eurocode 8 requires the analysis to account for the interaction between
the soil and the abutments, using realistic soil stiffness parameters. In that case, the bridge
may be designed for limited ductile behaviour with q 1.50. According to Part 2 of
Eurocode 8, the bridge may still be considered as locked-in (and designed with q 1.0 and
T 0 s) without the need to estimate the period T, if the abutments are embedded in stiff
natural soil over at least 80% of their total surface area that is in contact with the soil/ll
over the backll face of the abutments (Figure 4.12).

Although the seismic integrity and stability of bridges having a deck integral with the abutments
seems assured, their modelling and analysis are quite demanding, both for gravity loads and for
the design seismic action. It normally entails a fairly detailed nite-element discretisation of the
deck and the abutments and modelling of the soil behind the abutments with springs. Even
though these springs may be taken as linear for simplicity, the actual soilabutment interaction
is quite nonlinear and very different when the abutment moves towards the soil or pulls away.
Figure 4.12. A bridge considered as locked-in: the abutments are embedded in stiff natural soil over at
least 80% of their total surface area that is in contact with the soil over their backll face

Original
ground level

Original
ground level
hs2

ht2

Width b2

hs1
Width b1

ht1

61

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Part 2 of Eurocode 8 gives guidance for a conservative analysis, based on superposition of the
inertial response of the structure and of a simplied combination of limit equilibrium and an
elastic approach for the backll or the soil. For uncertain soil behaviour, it recommends using
upper- and lower-bound estimates of the soil stiffness. For single-span box-type culverts, it species as more realistic the use of an analysis based on kinematic compatibility of the structure and
the surrounding soil.
The Caltrans Seismic Design Criteria (Caltrans, 2006) give simple rules for the estimation of the
stiffness and the ultimate resistance of an abutment in the longitudinal direction of a bridge, due
to the passive resistance of a well-compacted backll. On the basis of static tests reported in
Maroney (1995) for a 1.7 m-tall abutment, the stiffness is about 7 MN/m per square metre
of the projection of the vertical face of the abutment onto a plane normal to the longitudinal
direction of the bridge. The ultimate passive resistance is reached when the top of the
abutment is pushed against the backll by about 2% of the abutment height (in units of MN
and m, this gives an ultimate resistance of about 0.14 times the projection of the vertical face
of the abutment onto a plane normal to the longitudinal direction, multiplied by the height
of the abutment). The results of a seismic analysis of the bridge using this longitudinal stiffness
for the pushed abutment are considered to apply, if the longitudinal displacement of the deck is
less than about 4% of the abutment height; if it exceeds that value by a factor of 2, the abutments are considered to contribute little against the longitudinal seismic action. In this latter
case, the analysis should be repeated with the abutment stiffness reduced to 10% of the value
quoted above. For longitudinal displacements between 2% and 4% of the abutment height,
linear interpolation between these extreme stiffness values may be used, and the analysis
repeated until convergence. Apart from a simple design tool to account for the abutment
backll interaction without recourse to the elaborate analysis alluded to in the previous
paragraph, the Caltrans (2006) simple rules may be used at the conceptual design phase to
estimate the contribution of relatively exible diaphragm abutments against the longitudinal
seismic action.
All things considered, the designers increased effort is a heavy price to pay for choosing this
robust bridge layout.

Clauses 2.3.6.3(2),
2.3.6.3(5), 6.6.1(2),
6.6.3.1(1), 6.6.3.1(2),
6.6.3.1(4) [2]

4.5.4
Deck on bearings
To avoid the main shortcoming of the integral connection, notably the restraint of longitudinal
deck movements and its consequences, a deck seated on the abutments through bearings can
move freely in the longitudinal direction (option 2a in Section 4.5.2). To this end, clearance
should be left between the main body of the deck and the abutment or its back-wall to
accommodate the sum of (see Section 6.8.2.1 of this Guide):
the full longitudinal displacement due to the design seismic action, plus
the displacement of the extreme bres of the deck end section towards the abutment (i.e.
the longitudinal displacement at the centroidal axis of the deck plus the rotation of the end
section multiplied by the distance of the extreme bres from the centroid) due to the quasipermanent gravity actions and prestressing, plus
3 50% of the end section displacement due to lengthening of the deck by its design thermal
action (the extreme one expected to take place during the bridge lifetime).
1
2

According to Part 2 of Eurocode 8, adequate clearance should be provided to protect important


components of the bridge (in this case the abutments). If such clearance is too large to be
provided at the roadway level, the movement joint there may be made wide enough to take
the sum of 2 and 3 above plus a fraction of 1; that is, of the deck longitudinal displacement
due to the design seismic action, implying that it will close when this fraction (recommended
to be 40% in Part 2 of Eurocode 8, see Section 6.8.2.2 of this Guide) is attained. Provisions
should then be made to limit the damage to the top of the abutment backwall, when the
movement joint closes at the roadway level. Figures 8.23 and 8.24 in Section 8.2.11.3 of this
Guide show an example of a practical means to this end: the deck slab protrudes from the
main body of the deck towards the backwall in order to reduce the gap at the roadway level,
while the top of the backwall is sacricial, to be knocked off without transmitting the forces
to the abutment and its foundation (this is option 2c of Section 4.5.2 turned into option 2a).
62

Chapter 4. Conceptual design of bridges for earthquake resistance

Seismic links are sometimes used to connect the deck to the abutments in the transverse direction.
They are arranged between the deck and an element (abutment or pier) supporting it at a
movable support joint, and link the two in a specic horizontal direction against the seismic
action. The link (usually in the form of shear keys or cables) is normally activated after a
specic horizontal displacement is attained and exhausts the gap of a shear key or the slack of
the cable (termed slack of the link). The design behaviour of seismic links according to Part 2
of Eurocode 8 is best understood by considering the following two limiting cases, depending on
the slack:
g
g

At the lower limit of zero slack, the link behaves as a xed bearing. The link is modelled in
the analysis as a xed constraint, and is capacity designed.
At the upper limit the slack is equal to the design seismic displacement (at the relevant
point and direction) of the bridge without the link. Then, it is meaningless to include the
link in the analysis model. The link plays the role of a second line of defence, against
seismic actions exceeding the design action. Part 2 of Eurocode 8 neither requires nor
recommends such a second line of defence. If, however, it is provided, the seismic design of
the bridge is not affected; the design of the link should be based on rational considerations
related by necessity to the reasons for its provision.

Part 2 of Eurocode 8 gives rules for the use of seismic links with slack between the above two
limiting cases:
(i) The link should be included in the analysis model, at least using the composite secant-toyield-point stiffness of the link and the supporting element (which is close to that of the
supporting element if the slack is small). Besides, measures to mitigate the impact are
required.
(ii) The design of the link and of the supporting element should be based on capacity
considerations.
If the abutment is extended upwards in the form of shear keys in contact with the sides of the
deck, the deck may be transversely restrained there (option 2b in Section 4.5.2). Normally,
shear keys are provided only at the sides of the deck, because internal ones (against the underdeck) are harder to inspect or maintain. An elastomeric bearing may be placed between each
shear key and the side of the deck to ensure full and soft contact under any seismic action and
to prevent local damage to the contact surfaces due to rotation of the ends of the deck in a
horizontal plane. In that case, under transverse seismic action the deck exes as a beam with
lateral elastic supports at the locations of the piers and the abutments. Note that the lateral
forcedeformation response of the system of the shear keys, the abutment with any wing-walls
and piles, and the soil is quite complex, and a composite elastic stiffness cannot be easily
estimated for it: its overestimation (let alone the assumption that it is laterally rigid) may be
unsafe for piers between the abutments. In addition, although shear keys are capacity
designed, if they do fail they will do so in a brittle fashion. Note also that, if the bridge is
designed for ductile behaviour in the transverse direction owing to a rigid transverse connection
with the piers, one may question whether the elasticperfectly plastic idealisation of the bridge
(the foundation of design for ductility) applies after the piers yield but the deck and its lateral
supports at the abutments are still elastic: in reality, the system will harden, at a hardening
ratio approximately equal to the elastic transverse stiffness of the deck on rigid or elastic end
supports divided by the elastic transverse stiffness of the pier system. For all these reasons, it
may be prudent to also analyse the bridge under transverse seismic action, disregarding the
shear keys and any restraint they offer, and to design the bridge for the envelope of design
action effects, with or without the shear keys.
In view of the questions raised in the previous paragraph, the designer may choose to let the deck
transversely free (option 2a in Section 4.5.2) and provide ample seating at the abutment to avoid
drop-off under the largest conceivable seismic action. To play it safe, shear keys may be provided
there as well, but with a clearance from the deck more than the transverse displacement due to the
design seismic action (unless the bridge is strongly curved, there is no transverse displacement due
to thermal or quasi-permanent actions). If, instead, the gap is narrower than this transverse
displacement (option 2c in Section 4.5.2), it will close during the design seismic action. Under
63

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

such circumstances, Part 2 of Eurocode 8 requires taking the deck as elastically supported on the
abutment, with a spring stiffness equal to the secant stiffness at yielding of the shear key after
closure of the gap (i.e. a stiffness equal to the yield force of the shear key divided by the clearance
plus the elastic deformation of the shear key until yielding). This is not very convenient, as the
shear key has not been dimensioned at this stage of the design. So, option 2c is not a very practical design alternative for the transverse direction.

4.6.

The foundations

Bridges are nowadays built on spread footings or on pile foundations, less commonly on deep
caissons.
Bridges on spread foundations are supported close to the ground surface on rm soil layers or
rock, and have performed well in earthquakes. Spread foundations directly transmit loads
from the superstructure to competent ground. They are dened as such foundations if they
have a ratio of the embedment depth to the foundation width of not more than 0.5. Deeper foundations are referred to as caisson foundations. Deep caisson foundations of bridge piers are
today normally used wherever the available plan area for the foundation is severely limited,
not allowing the use of a spread or pile group foundation.
On a site with weak upper soil layers, the bridge is supported on deep foundations transferring
the vertical and lateral forces to stronger soil layers beneath the soft material. Bridges on soft
clay, silt or loose saturated sand have been damaged by the amplication of the ground
motion or soil failure in earthquakes. Unless massive soil failure has occurred, pile foundations
have performed well in past earthquakes, even when other bridge elements have suffered considerable damage. By contrast, bridges on liqueable soil deposits or soft sensitive clays have
been particularly vulnerable to earthquakes: soil liquefaction can cause a loss of bearing
capacity and, sometimes, lateral movement of the substructure.
Clause 5.4.2(1) [3]

The seismic response of pile foundations to strong earthquake shaking is very complex. It is
controlled by inertial interaction between the superstructure and the pile foundation, kinematic
interaction between foundation soils and piles, and the nonlinear stressstrain behaviour of soils
and of the soilpile interface. In addition, at some sites the build-up of high pore water pressures
during the earthquake or liquefaction add to the complexity. Many different materials and
geometries have been used for pile designs. Although numerous examples of long-span bridges
in seismic areas founded on timber piles still exist in California, the current trend is to use
concrete piles (reinforced or prestressed) or steel piles (H sections, shells or concrete-lled
shells). A special case is an integral pileshaft column arrangement, where the pile is not
connected to a pile cap but extended in the superstructure as a column.

Clause 2.2.4.2 [1]


Clause 5.2(1) [3]
Clauses 5.1(1), 5.3.2
[3]

Unlike other types of structures for which only one foundation type may in general be used,
different foundation types may be encountered in the same bridge.
The basic principles of foundation design require the foundation be able to safely transfer to the
ground the applied loads. Accordingly, it should be mechanically stable and prevent detrimental
displacements. Soilstructure interaction should be assessed where necessary, also taking into
account the relevant provisions in Part 5 of Eurocode 8 (CEN, 2004). To ensure stability, the
foundation must possess the required factors of safety against bearing, sliding and overturning
failure mechanisms. The relevance of these failure modes to each foundation type is shown in
Table 4.2. In addition, the structural elements of the foundation are designed to resist the
action effects.
Table 4.2. Foundation failure mechanisms and stability verications

64

Foundation type

Bearing capacity

Overturning

Sliding

Horizontal displacement

Spread foundation
Caissons
Piles

Yes
Yes
Yes

Yes

Yes

Yes
Yes

Chapter 4. Conceptual design of bridges for earthquake resistance

Because it is difcult to inspect or repair foundations after an earthquake, it is a common practice to


restrict damage to the foundation to a minimum, so that operation of the bridge can easily restart
without repair of the foundations. In general, bridge foundations should not be intentionally used
as sources of hysteretic energy dissipation, and therefore should, as far as practicable, be designed
to remain elastic under the design seismic action. To this end, if the bridge is designed for limited
ductile behaviour, its foundation is designed to remain elastic. If the bridge is designed for ductile
behaviour, the foundation is designed to resist the overstrength capacity of the pier column(s) or
wall for plastic hinges presumed to form at their base. However, in some instances such a design is
not possible. A prime example is a bridge structure with a large force capacity due to factors other
than the earthquake. Under such circumstances, piles are allowed to develop a plastic hinge next
to their connection to the pile cap, where large bending moments develop. The head of the pile up
to a distance to the underside of the pile cap of three times the pile cross-sectional dimension, d, is
detailed as a potential plastic hinge region. To this end, it should be provided with transverse and
connement reinforcement following the rules for pier columns designed for ductile behaviour.
The same applies to regions of the pile up to a distance of 2d on each side of an interface between
two soil layers with a large shear stiffness contrast (ratio of shear moduli greater than 6.0).

Clauses 5.8.1(1),
6.4.1(1), 6.4.2(1)
6.4.2(4) [2]
Clause 5.4.2(7) [3]

Inclined piles are usually not recommended for transmitting lateral loads to the soil. If used, they
should be designed to safely carry the axial loads as well as bending moments arising from soil
settlements. One additional reason for not favouring inclined piles is their less ductile behaviour
as opposed to vertical piles subjected to pure bending.

Clause 5.4.2(1) [3]

Piles required to resist tensile forces or assumed as rotationally xed at the top should be
provided with anchorage in the pile cap to enable the development of the pile design uplift resistance in the soil, or of the design tensile strength of the pile reinforcement, whichever is lower. If
the part of such piles embedded in the pile cap is cast before the pile cap, dowels should be
provided at the interface of the connection.

Clause 5.8.4(3) [1]

REFERENCES

Bardakis V (2007) Displacement-based seismic design of concrete bridges. Doctoral thesis,


Department of Civil Engineering, University of Patras.
Caltrans (2006) Seismic Design Criteria, version 1.4. California Department of Transportation,
Sacramento, CA.
CEN (Comite Europeen de Normalisation) (2000) EN 1337-2:2000: Structural bearings Part 2:
Sliding elements. CEN, Brussels.
CEN (2004) EN 1998-5:2004: Eurocode 8 Design of structures for earthquake resistance Part 5:
Foundations, retaining structures, geotechnical aspects. CEN, Brussels.
CEN (2005) EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance Part 2:
Bridges. CEN, Brussels.
CEN (2008) CEN/TC 1992-4-1:2009: Design of fastenings for use in concrete. CEN, Brussels.
b (2000) Guidance for Good Bridge Design. b Bulletin 9. Federation Internationale du Beton,
Lausanne.
b (2004) Precast Concrete Bridges. b Bulletin 29. Federation Internationale du Beton, Lausanne.
b (2007) Seismic Bridge Design and Retrot Structural Solutions. b Bulletin 39. Federation
Internationale du Beton, Lausanne.
b (2009) Structural Concrete Textbook on Behaviour, Design and Performance, vol. 1, 2nd edn.
b Bulletin 51. Federation Internationale du Beton, Lausanne.
b (2011) Design of Anchorages in Concrete. b Bulletin 58. Federation Internationale du Beton,
Lausanne.
b (2012) Model Code 2010 Final Draft, vol. 2. b Bulletin 66. Federation Internationale du
Beton, Lausanne.
Maroney BH (1995) Large scale bridge abutment tests to determine stiffness and ultimate strength
under seismic loading. PhD thesis, University of California, Davis, CA.
Priestley MJ, Seible F, Calvi GM (1996) Seismic Design and Retrot of Bridges. Wiley-Interscience,
New York.
Stathopoulos S, Kotsanopoulos P, Stathopoulos E et al. (2004) Votonosi bridge in Greece.
Proceedings of the b Symposium: Segmental Construction, Delhi.

65

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.067

Chapter 5

Modelling and analysis of bridges for


seismic design
5.1.

Introduction: methods of analysis in Eurocode 8

An analysis carried out in the framework of seismic design determines by calculation the effects
of the design seismic actions in terms of internal forces and deformations, to be used for the
dimensioning of its members; that is, to verify their cross-sectional size and in the case of
concrete members to determine the amount and location of their reinforcement.
This chapter is limited to the essentials for the application of well-established analysis methods to
the design of bridges for earthquake resistance according to Eurocode 8. The reader is assumed to
be conversant with the fundamentals of structural dynamics and their application for seismic
analysis.
The seismic design of bridges according to Eurocode 8 is primarily force-based. Its main workhorse is linear-elastic analysis based on the design response spectrum of Section 5.3; that is, on
the 5%-damped elastic spectrum divided by the behaviour factor q, which accounts on one hand
for ductility and energy dissipation capacity and on the other for overstrength. Its counterpart in
the USA is the force reduction factor or response modication factor, R.
Eurocode 8 provides two alternative methods for linear-elastic seismic analysis:

Clauses 4.2.1, 4.2.2 [2]

(a) Linear static analysis (termed the fundamental mode method in Part 2 of Eurocode 8
(CEN, 2005a) and the lateral force method in Part 1 (CEN, 2004a), also known in
practice as equivalent static analysis).
(b) Modal response spectrum analysis, as known in practice and called in Part 1 of Eurocode
8, termed in Part 2 just the response spectrum or linear dynamic analysis or analysis
with a full dynamic model.
Eurocode 8 adopts analysis method (b) as the reference method for the design of bridges, and
fully respects its rules and results. It allows its application to any bridge, except to those with
a strongly nonlinear seismic isolation system. If both methods of linear-elastic seismic analysis
are applicable for the design of a given bridge, an analysis of type (b) gives on average a more
even distribution of peak internal forces over the bridge, translated to material savings. If its
results are used for member dimensioning, the overall inelastic performance of the bridge may
be expected to be better, because peak inelastic deformations are normally closer to its predictions than to those of an analysis of type (a). So, as reliable and efcient computer programs
for modal response spectrum analysis in 3D are widely available nowadays, it can be adopted
as the single analysis tool for the seismic design of bridges. It is worth noting, however, that
quite a few bridges are close to a single-degree-of-freedom (SDoF) system; hence, linear static
analysis based on the fundamental mode provides considerable insight into their seismic
response and still has a role as a practical seismic design tool for bridges.
Eurocode 8 recognises an important role for:

Clauses 4.1.6(1),
4.2.1.1(1) [2]

Clauses 4.1.9(2), 4.2.4,


4.2.5 [2]

(i) nonlinear static analysis (commonly known as pushover analysis)


(ii) nonlinear dynamic (time-history or response-history) analysis
67

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

in the seismic design of bridges. However, their stand-alone application is limited to the design of
irregular ductile bridges (see Section 5.10.2 of this Guide) with deformation-based verication
of their ductile members and to bridges with seismic isolation. The limitation is due to concerns
about the correct and prudent application by practitioners of these more rational methods, which
have a very different safety format and entail analysis of structures with known member dimensions and reinforcement in the same phase of the design process as their verication.
Clause 4.2.3.1 [2]

Unlike Part 1 of Eurocode 8, which does not explicitly mention linear time-history analysis, Part
2 of Eurocode 8 does mention the method. However, it is not a very practical alternative to linear
modal response spectrum analysis.

Clauses 2.3.1.1(5)
When linear analysis with the design response spectrum is applied, internal forces for the dimen2.3.1.1(8), 2.3.3, 2.3.4, sioning are taken as equal to those estimated from it, except for regions or members of ductile
4.2.4.4(2) [2]
bridges intended to remain elastic, which are dimensioned on the basis of capacity design

actions. Displacements due to the seismic action are obtained from those derived from the
linear analysis after corrections for (a) any difference of the effective stiffness values from
those initially assumed, (b) damping other than the default value of 5% and (c) deviations of
short-period bridges from the equal displacement rule (see Sections 2.3.2.2, 5.8.4 and 5.9.1).
By contrast, when nonlinear analysis is applied, all seismic action effects (internal forces,
displacements and deformations) are taken as equal to those derived from it. Only safety
factors intervene between these demands and the capacities against which they are checked.

5.2.

The three components of the seismic action in the analysis

Clause 3.2.3.1.1(2) [1] The three components of the seismic action are taken to act concurrently on the bridge.
Clauses 3.1.2(1),
Time-history analysis, be it linear elastic or nonlinear, is indeed carried out applying
3.1.2(2), 3.2.3(5) [2]
simultaneously all seismic action components of interest (the two horizontal ones and occasion-

ally the vertical).


Clause 4.3.3.5.1(2) [1] Linear static or modal response spectrum analyses give just the peak values of each seismic action
Clauses 4.2.1.4(1),
effect of interest due to the individual seismic action components. These peak values are statisti4.2.2.1(3) [2]
cally combined as outlined below. They are denoted here by EX and EY for the two horizontal

components and EZ for the vertical. As they do not occur simultaneously, a combination rule
of the type: E EX EY EZ is too conservative. The reference rule in Part 2 of Eurocode 8
is the square root of the sum of squares (SRSS) combination of EX , EY , EZ in Smebby and
Kiureghian (1985):
E

q
EX2 EY2 EZ2

D5:1

Clause 4.3.3.5.2(4) [1] Part 2 of Eurocode 8 also accepts as an alternative a linear combination rule of the type:
Clause 4.2.1.4(2) [2]

E |EX| l|EY| l|EZ|

(D5.2a)

E l|EX| |EY| l|EZ|

(D5.2b)

E l|EX| l|EY| |EZ|

(D5.2c)

where the meaning of is superposition. A value l  0.275 provides the best average agreement with the result of Eq. (D5.1) in the entire range of possible values of EX , EY , EZ . In
Eurocode 8 this optimal l value has been rounded up to l 0.3, which may underestimate
the outcome of Eq. (D5.1) by at most 9% (when EX , EY and EZ are about equal) and may overestimate it by not more than 8% (when two of these three seismic action effects are an order of
magnitude less than the third).
Clauses 3.1.2(1),
4.1.1(2) [2]

68

To combine the effects of the seismic action components as outlined above, it is computationally
convenient to do the analysis on a 3D model of the bridge, with all seismic action components of
interest applied to it separately but their effects combined within the same computational
environment and analysis run. So, it is not convenient to use a separate model for the
response to the horizontal component in the longitudinal direction of the bridge, another one
in the transverse and a third for the vertical component (if of interest), as allowed by

Chapter 5. Modelling and analysis of bridges for seismic design

Eurocode 8. However, as the seismic response to each one of these components is distinctly
different, it is intuitively appealing and very didactic to consider them separately using a different
model for each one. This is often done in this chapter.
The horizontal components of the seismic action are most often taken parallel to the longitudinal and the transverse direction of the bridge. If the deck is straight and the abutments
are at right angles to its axis, the denition of these directions is clear. If the deck is curved
but fairly long, the longitudinal direction may be taken as that of the chord connecting the
two points where the deck axis intersects the line of support at each abutment; the transverse
direction is at right angles to it. If the bridge is skewed and it has a fairly wide deck, it may
make more sense to dene the longitudinal and transverse directions as normal and parallel
to the abutments, because these are the directions in which the bridge, the abutments and the
piers (normally aligned parallel to the abutments) primarily work. If bearings or other devices
(e.g. shear keys) restrain the deck in one direction but not in the other (where it may be
supported with a certain horizontal exibility or be free to move), these two directions also
lend themselves as longitudinal and transverse. A very important point is that the behaviour
factor, q, may be taken different in these two directions, depending on the way the piers are
connected to the deck and the ratio of their shear span (moment-to-shear ratio) to their crosssectional depth in that direction. If these features are radically different along two orthogonal
directions, these directions are prime candidates for the horizontal seismic action components
as well.

5.3.

Design spectrum for elastic analysis

For the horizontal components of the seismic action the design response spectrum to be used in
linear elastic analysis is given by different expression in four different period ranges (repeated
below from Section 3.1.3 for reasons of their importance and of convenience):

Clauses 2.1(2),
3.2.4(1), 4.1.6(1) [2]
Clause 3.2.2.5(4) [1]

Short-period range:




2 T 2:5 2
0  T  TB : Sa;d T ag S

3 TB q
3

D5:3a

Constant spectral pseudo-acceleration range:


TB  T  TC : Sa;d T ag S

2:5
q

D5:3b

Constant spectral pseudo-velocity range:


TC  T  TD : Sa;d T ag S

 
2:5 TC
 bag
q T

D5:3c

Constant spectral displacement range:


TD  T: Sa;d T ag S



2:5 TC TD
 bag
q
T2

D5:3d

The design spectrum in the vertical direction has also been described in Section 3.1.3 of this
Guide.

5.4.

Behaviour factors for the analysis

The concept and role of the behaviour factor q has been outlined in Section 2.3.2.2 of this Guide
as far as the design of ductile bridges is concerned and in Section 2.3.3 for design for limited
ductile behaviour. Section 5.3 illustrated the use of the behaviour factor q in the design
response spectrum. Before venturing into the rules and details of analysis and modelling for
the purposes thereof, the values of the behaviour factor q specied in Part 2 of Eurocode 8 for
the design of these two different types of bridges for the horizontal components of the seismic
action are given. Note that the value of q enters in the design response spectrum and therefore
should be known before any analysis can be carried out.
69

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Clauses 4.1.6(3),
4.1.6(12) [2]

As the design response spectrum is specied separately for each component of the seismic action,
the q factor can be different for each of them. and normally is. Starting with the vertical direction,
bridge piers and bearings cannot develop ductile behaviour in pumping concentric axial compression and tension, at least at the short natural periods of the vertical eigenmodes. Additionally, vertical modes of vibration excite mainly bending of the deck in a vertical plane; but the deck
is meant to remain elastic under the design seismic action and beyond (except in exible and
ductile continuity top slabs between adjacent simply supported spans). So, we use q 1 in the
vertical direction.
The value of q in the two horizontal directions depends on how the bridge is congured to
accommodate the horizontal seismic displacements of the deck in each of them. Sections 2.3.1
and 4.1 of this Guide highlighted three options to accommodate these horizontal displacements:
in exible bearing-type devices arranged over an effectively horizontal interface between
the deck and the abutments and piers
2 in exural plastic hinges at the base and possibly at the top of piers rigidly connected
with the deck
3 at the interface between the foundation element of the piers and the ground or in plastic
hinges in foundation piles
1

as well as the option of:


4

locking the bridge in the ground, by connecting the deck and the abutments into an
integral system that follows the ground motion with little additional deformation of its
own.

Clauses 4.1.6(3),
4.1.6(12) [2]

As explained in Section 2.3.2.5, neither option 1 nor option 3 with the base of the piers sliding
with respect to the soil lend themselves to the development of signicant inelastic deformations
in ductile members, such as the piers. Seismic design using these options should therefore be
elastic, with q 1. For seismic design with a q factor that is greater than 1.0 owing to ductility
and the energy dissipation capacity we are left with options 2 and 4 (and with option 3 for
plastic hinging in foundation piles).

Clauses 4.1.6(3),
4.1.6(7) [2]
Clauses 4.1.6(9),
4.1.6(10), 6.7.3(4),
6.7.3(9) [2]

The maximum values of q specied in Eurocode 8 for these options are given in Table 5.1.
Locked-in bridges, or dynamically independent parts thereof, are considered to follow the horizontal motion of the ground without appreciably amplifying it. Their design spectral acceleration
may be taken as equal to the design peak ground acceleration (i.e. as for T 0 and q 1). This
category includes abutments connected to the deck through movable bearings, or integral bridges
having rigid horizontal connection of the deck to the abutments and a fundamental period T in
the horizontal direction of interest less than 0.03 s. This case corresponds to the last row of
Table 5.1. The row above it refers to bridges with an essentially horizontal deck rigidly connected
to both abutments (either monolithically or through xed bearings). For such bridges, Eurocode
8 requires accounting for the interaction between the soil and the abutments, using realistic soil
stiffness parameters. If T is longer than 0.03 s, the design response spectrum should be entered
with that value of T and with q 1.5. Estimation of T may be avoided and the bridge considered
as locked in if its abutments are embedded in stiff natural soil over at least 80% of their surface
area that is in contact with the soil/ll over the backll face of the abutments (see Figure 4.12).

Clause 4.1.6(6) [2]

As already noted in Section 2.3.1, if plastic hinges can form in an inaccessible part of any pier,
the design of the bridge should be based on just 60% of the value of q in Table 5.1 (but not less
than q 1). Part 2 of Eurocode 8 considers as accessible the base of piers deep in earth ll, but
as inaccessible those in deep water or groundwater, or piles under large pile caps. At any rate,
locations of the pier below the normal water table at a depth that can be accessed and drained
with reasonable retaining and pumping effort and cost may be considered as accessible for
repairs.

Clause 4.1.6(5) [2]

The values of q given in Table 5.1 for reinforced concrete piers apply as long as the maximum
value of the axial force ratio hk (axial load due to the design seismic action in the horizontal

70

Chapter 5. Modelling and analysis of bridges for seismic design

Table 5.1. Behaviour factor q for use with the horizontal components of the seismic action
Type of ductile member

Seismic behaviour
Limited ductile

Ductile

Reinforced concrete piers


Vertical piers in bending
Inclined struts in bending

1.5
1.2

3.5a
2.1a

Piles under pile caps


Vertical piles in bending
Inclined piles

1.0
1.0

2.1
1.5

Steel piers
Vertical piers in bending
Inclined struts in bending
Piers with normal bracing
Piers with eccentric bracing

1.5
1.2
1.5

3.5
2.0
2.5
3.5

Abutments rigidly connected to the deck


In general
Locked-in bridges

1.5
1.0

1.5
1.0

Arches

1.2

2.0

a
If the minimum among all piers of the pier shear span ratio, Ls /h (where Ls M/V is the distance from the base of
the pier to the inexion point for the seismic action direction considered and h ispthe
cross-section
depth in that

direction), is less than 3.0, these value of the behaviour factor are multiplied by Ls =3h. For piers with sides skewed
to the seismic action component considered, use the minimum value of Ls /h along the two sides

direction of interest and the concurrent gravity loads, Nd , normalised to the product of the pier
section area and the characteristic concrete strength, Ac fck) among all piers does not exceed 0.3.
As already noted in Section 2.3.3, if this largest value of hk Nd/Ac fck exceeds 0.6, the q factor of
the bridge is taken as equal to 1.0. For a value of hk between 0.3 and 0.6, linear interpolation
between 1.0 and the value in Table 5.1 is allowed. Again, it is the largest value of hk among
all reinforced concrete piers that controls the q factor of the entire bridge. It should be noted,
though, that in normal practice the values of hk are fairly low. If nothing else, the upper limits
set by Eurocode 2 on the slenderness of pier columns to avoid lengthy and cumbersome calculations of second-order effects under factored gravity loads (the persistent and transient
design situation of EN 1990) lead to fairly large-sized pier columns and drive down their hk
values (see Section 4.4.3.1 of this Guide).
The shear span of the pier, Ls M/V, in the longitudinal and the transverse directions of the
bridge needs to be estimated before any analysis for the seismic action yields the values of the
moment, M, and shear, V, at the sections where plastic hinges may form. The values of Ls
quoted in Section 2.3.2.3 in cases 1 to 4 may be conveniently used for that purpose. Case 1
therein (namely that of seismic response in the longitudinal direction with the top of the pier
columns monolithically connected to the deck) may be rened as follows.
Let us denote by EId the elastic rigidity of the deck for bending in a vertical plane through
its longitudinal axis. For bending within that same plane, EIp is the total effective rigidity of
the pier, which may consist of n  1 columns, each with an effective rigidity EIn ; then, EIp
Sn(EI )c (see Section 5.8 for the effective rigidities). If Ld is the average span length on either
side of the pier (with an exterior span freely supported at the abutment taken with twice its
span value in this averaging) and Hp the clear pier height, the relative stiffness of the deck to
the pier is dened as
k

EId Hp
EIp Ld

D5:4a
71

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Then, the shear span at the base of the pier is


Ls

k 1=6 Hp
k 1=12 2

D5:4b

Case 4 in Section 2.3.2.3 (namely that of single-column piers as vertical cantilevers for the
response in the transverse direction) may be rened as follows, to include the effects of (a) the
tributary rotational mass moment of inertia, Iu,d , of the deck in a vertical plane through the transverse direction and (b) the vertical distance between the pier top and the point of application of
the deck inertia force, taken for convenience at the centroid of the deck section, at a distance ycg
from the bottom of the deck:
Ls Hp ycg

1:5I
 u;d 
Md Hp ycg

D5:5

In Eq. (D5.5) rm,d is the radius of gyration of the deck mass (square root of the ratio of Iu,d to
the tributary deck mass, Md). Both Md and Iu,d refer to the full average span length Ld . For
completeness and future reference, they are given here as




c 2 qk
g
rsurf tsurf bsurf side
D5:6a
Md Ld rc Ac;d
g
g
"
!
!#


c2 qk
b2surf
gside b2side
2
2
rsurf tsurf bsurf
hd  ycg
hd  ycg
Iu;d Ld rc Jp;d
g
12
g
4
D5:6b
In Eqs (D5.6) rc and rsurf are the mass densities of concrete and of the surfacing material (or of
the ballast, in railway bridges), respectively; Ac,d and Jp,d are the surface area and the polar
moment of inertia of the deck section, respectively; c2qk is the quasi-permanent value of the
uniform trafc load; bsurf is the width of the deck over which it is applied; tsurf is the thickness
of the surfacing (or ballast) considered to be applied over bsurf as well; gside is the total weight
of hand rails, parapets and kerbs at the two sides of the deck per linear metre of deck and
bside their distance across the deck; hd is the deck depth; and g 9.81 m/s2, the acceleration of
gravity.
Part 2 of Eurocode 8 recommends taking c2 0 for bridges with normal trafc and footbridges.
For road bridges with severe trafc conditions (dened in a note as motorways and other roads
of national importance) it recommends c2 0.2, and for railway bridges with severe trafc
conditions (dened in a note as intercity rail links and high-speed railways) c2 0.3. In both
cases, only the characteristic value, qk , of the uniform trafc load (UDL) of Load Model 1
(LM1) with all of its nationally applying adjustments is recommended to contribute to the
quasi-permanent part of the severe trafc on the bridge. The quasi-permanent value of the
trafc loads is not only considered to give together with the permanent actions the masses
that are subjected to the seismic action and produce the inertia forces but is also combined
with them and the design seismic action in the seismic design situation for which the performance requirements are veried (see Section 6.2 of this Guide).
The discussion above tacitly implies that the bridge deck is (effectively) straight and that the principal directions of bending of the pier sections are parallel and orthogonal to the deck and, hence,
to the longitudinal and the transverse direction of the bridge. The footnote in Table 5.1 mentions
piers with sides skewed to these directions; for example, wall-like or hollow-rectangular piers
with sides parallel and orthogonal to abutments that are skewed to the axis of the deck. Such
piers present to the longitudinal and transverse directions of the bridge their effective rigidities
in these two directions, as well as a coupling (cross-)rigidity. In terms of moments of inertia,
they work with moments of inertia IL , IT and ILT (cross-moment of inertia) resulting from
rotation of the principal ones, I1 and I2 , to the directions parallel and orthogonal to the longitudinal direction. For bending in a vertical plane in the transverse direction of the bridge, the value of
IT applies. Such bending involves both depths of the section parallel to its principal directions (i.e.
both sides). The larger of these two depths vis-a`-vis the corresponding shear span controls the
72

Chapter 5. Modelling and analysis of bridges for seismic design

ductility of the pier (or lack of it). So, it is the smallest of the two shear span ratios that determines
the value of q for both directions of the seismic action.

5.5.

Modal response spectrum analysis

5.5.1
Modelling
5.5.1.1 Introduction
Guidance for modelling given below while presenting the modal response spectrum method of
analysis also applies to the other analysis methods highlighted in Sections 5.6 and 5.10. By
their nature as simplications of modal response spectrum analysis, the linear methods in
Section 5.6 are also open to simplied modelling. By contrast, the nonlinear methods of
Section 5.10 normally require extension of certain modelling aspects into the nonlinear regime.
Linear analysis methods, such as those in Sections 5.5 and 5.6, are not meant to account for
nonlinear behaviour other than in locations intended for energy dissipation through ductility,
notably in exural plastic hinges. Even this type of nonlinearity is dealt with in a conventional,
approximate way, via the behaviour factor, q. The effects of other types, even when they are
of minor importance for the global response, cannot be properly captured by linear analysis.
This includes friction in sliding bearings, abrupt engagement of shear keys when the gap with
the deck closes, the interaction between the backll and the abutment and its wing walls,
closing and re-opening of deck movement joints at the abutments or in between (owing not
only to the longitudinal response but to the transverse response as well, which may close a
joint just on one side of the deck), etc. Such phenomena should be dealt with by nonlinear
analysis. As pointed out in Chapter 4, if the designer is not prepared to resort to nonlinear
analysis, they should avoid altogether sliders, intermediate movement joints, gaps between the
deck and shear keys or the abutments that close in the seismic design situation, or engineer
their way around the nonlinearity. For example, the information from Caltrans (2006)
highlighted in the last paragraph of Section 4.5.3 can be used to design a sacricial backwall
that will be knocked off when the movement joint at roadway level closes in the seismic design
situation.
5.5.1.2 Modelling of the deck and the piers
Clauses 4.1.1(1),
The model of a bridge for the purposes of a modal response spectrum analysis (or a full
dynamic model in the language of Part 2 of Eurocode 8) should account for the distribution 4.1.2(2), 4.2.1.1(2) [2]
of mass all over the bridge; that is, all over the deck and in the piers, down to the top of competent ground. Regarding the mass of the piers, the universal 10% rule of thumb of structural
design implies that the mass of a pier should be accounted for if it is more than 10% of the
tributary mass of the deck. At any rate, it costs nothing to place a string of approximately
equidistant intermediate nodes along the centroidal axis of a pier and lump there its continuously distributed mass. Piers are normally 1D elements, with a straight axis and a constant
or varying (tapered or ared) section. Prismatic 3D beam/column (sub)elements, with
constant cross-sectional properties along their length, connect adjacent intermediate nodes of
the pier. If the pier section varies, the spacing of intermediate nodes should be sufcient for
a good stepwise approximation of this variation. If it does not, at least three intermediate

Figure 5.1. Discretisation of a concrete bridge with a box girder deck built with the balanced cantilever
method, showing the sections at all nodes of the piers and the deck

73

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

nodes are placed. Similar is the model used for decks consisting of a single girder (commonly a
box girder, of concrete, steel or composite steel and concrete). Figure 5.1 depicts an example
of such a model for a concrete bridge with monolithic connection of the piers with the box
girder deck.
In a discretisation with prismatic 3D beam/column elements, each node possesses all six degrees of
freedom (DoFs): three translations and three rotations (certain DoFs of support nodes are of course
restrained). Masses are assigned to all three translational DoFs of each deck node. Their values are
given by Eq. (D5.6a), with the span length, Ld , replaced by the average length of the two prismatic
deck elements on either side of the node. A rotational mass moment of inertia, Iux , should always be
assigned to the rotational DoF about the centroidal axis, x, of the deck. Its value is given by Eq.
(D5.6b), with the span length, Ld , again replaced by the average length of the two prismatic deck
elements on each side of the node. Similar is the lumping of the mass and of the rotational mass
moment of inertia, Iux , of the piers along their longitudinal axis, x. As the nodes are often closely
spaced along the longitudinal axis of the deck or the pier, there is no need to assign rotational
mass moments of inertia to the two other rotational DoFs, namely those about the centroidal
axes y and z of the section. Closely spaced nodes along the deck and the pier give a model resembling
the continuous spread of the mass all along the deck and the piers, and allow determining modal
shapes that reect this distribution. For example, higher modes may induce more than one inexion
point between the top and bottom of a pier or between adjacent joints of the deck and piers. These
modes cannot be captured, unless a good number of nodes are used between adjacent physical joints
of the pier and the deck. Figures 5.2(d) and 5.3(d)5.3(n), as well as Figure 8.6 in Section 8.2.6 of this
Guide, show examples of such modes for bridges with monolithic connection of the piers and the
box girder deck. Figure 5.3 is for a bridge having very dissimilar pier heights, with the upper
30 m of all piers in the form of twin blades, and the lower part if any with a hollow rectangular
section. Figures 8.32 to 8.35 in Section 8.3.3.4 depict also the important modes of the bridge model
of Figure 8.27.
Masses should be assigned to all nodes down to the top of ground that is competent enough to be
included in the model with its stiffness (be it innite). Liqueable and very weak cohesive or silty
soils are excluded from being considered as competent. Nodes on piles (or other foundation
elements) below the top of competent ground may be considered as massless. By the same
token, earth or water pressures should not be considered to act on these nodes.
A fairly rened model of the deck is normally necessary for the analysis for gravity loads
(permanent and due to trafc, often of quite non-uniform distribution owing to wheel or lane
loads), in order to dimension the deck for the ultimate limit state (ULS) and the serviceability
limit state (SLS) in out-of-plane bending due to the relevant combinations of these actions
(including intermediate stages of construction and taking into account the redistribution of
action effects due to creep, in case the deck consists partly or fully of concrete). If the same
discretisation is used for the modal response spectrum analysis, its results are conveniently
combined with those of the analysis for the permanent actions plus the quasi-permanent value
of the trafc loads (after redistribution due to creep, etc.), to give the design action effects in
the seismic design situation. Although convenient, the choice of the same deck model for the
seismic analysis is not necessary, as inertial seismic loads, being proportional to mass
multiplied by response acceleration, are fairly uniformly distributed over the deck surface.
Therefore, for the purposes of the global seismic analysis, the deck may be modelled as a
spine of beam elements. Such a model cannot capture the intricate distribution of seismic
action effects at and around monolithic connections of the deck with the piers especially
multi-column ones. If the detailed distribution of seismic action effects in these regions is of
interest, it may be derived from those from the global seismic analysis by subjecting the
detailed model of the deck used in the gravity load analyses to unit static forces or moments
in the relevant directions coming from the pier column(s). Note also that in bridges designed
for ductile behaviour the deck and its monolithic connections to pier columns are veried for
capacity design effects.
If the deck consists of a concrete slab, its analysis for gravity loads is normally based on a
very rened mesh of shell-type nite elements (i.e. a combination of plate nite elements
for out-of-plane bending, and plane-stress ones for the in-plane loads and response).
74

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.2. Shapes, periods and participation factors of four modes of a bridge of the type depicted in
Figure 5.1, which capture approximately 90% of the total mass in each of two horizontal directions: (a, c,
d) transverse; (b) longitudinal direction (Bardakis, 2007)

Mode 1
Axis
Participation factors
Effective modal mass: %

Period: 2.0483 s
X
Y
Z
0.00
0.00 113.32
0.00
0.00
73.72

(a)

Mode 2
Axis
Participation factors
Effective modal mass: %

Period: 1.8889 s
X
Y
Z
127.04 1.12
0.00
91.96 0.01
0.00

Mode 5
Axis
Participation factors
Effective modal mass: %

Period: 0.6810 s
X
Y
Z
0.00
0.00
46.20
0.00
0.00
12.26

Mode 16
Axis
Participation factors
Effective modal mass: %

Period: 0.1532 s
X
Y
Z
0.00
0.00
24.22
0.00
0.00
3.37

(b)

(c)

Y
X
Z
(d)

75

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 5.3. Shapes, periods and participation factors of 14 modes capturing approximately 90% of the
total mass of the bridge depicted in Figure 4.6 in (a, d, g, i) the longitudinal direction and (b, c, e, f, h,
jn) the transverse direction (Bardakis, 2007)

Mode 1
Axis
Participation factors
Effective modal mass: %

Period: 2.6057 s
X
Y
Z
106.50 0.05
0.00
72.23 0.00
0.00

Mode 2
Axis
Participation factors
Effective modal mass: %

Period: 1.7494 s
X
Y
Z
0.00
0.00 98.08
0.00
0.00 61.95

Mode 4
Axis
Participation factors
Effective modal mass: %

Period: 0.5015 s
X
Y
Z
0.00
0.00 39.92
0.00
0.00 10.26

Mode 6
Axis
Participation factors
Effective modal mass: %

Period: 0.4863 s
X
Y
Z
45.46 0.76 39.92
13.16 0.00 10.26

(a)

(b)

(c)

(d)

76

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.3. Continued

Mode 7
Axis
Participation factors
Effective modal mass: %

Period: 0.4717 s
X
Y
Z
0.00
0.00 28.13
0.00
0.00
5.09

Mode 8
Axis
Participation factors
Effective modal mass: %

Period: 0.3890 s
X
Y
Z
0.00
0.00 16.18
0.00
0.00
1.69

Mode 16
Axis
Participation factors
Effective modal mass: %

Period: 0.2580 s
X
Y
Z
21.29 3.35
0.00
2.89 0.07
0.00

Mode 17
Axis
Participation factors
Effective modal mass: %

Period: 0.2537 s
X
Y
Z
0.00
0.00 18.96
0.00
0.00
2.31

(e)

(f)

(g)

(h)

77

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 5.3. Continued

Mode 22
Axis
Participation factors
Effective modal mass: %

Period: 0.2305 s
X
Y
Z
14.11 4.79
0.00
1.27 0.15
0.00

Mode 27
Axis
Participation factors
Effective modal mass: %

Period: 0.1461 s
X
Y
Z
0.00
0.00 15.28
0.00
0.00
1.50

Mode 28
Axis
Participation factors
Effective modal mass: %

Period: 0.1408 s
X
Y
Z
0.00
0.00 18.84
0.00
0.00
2.29

Mode 30
Axis
Participation factors
Effective modal mass: %

Period: 0.1209 s
X
Y
Z
0.00
0.00 17.50
0.00
0.00
1.97

(i)

( j)

(k)

(l)

78

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.3. Continued

Mode 31
Axis
Participation factors
Effective modal mass: %

Period: 0.1177 s
X
Y
Z
0.00
0.00 16.35
0.00
0.00
1.72

Mode 49
Axis
Participation factors
Effective modal mass: %

Period: 0.0749 s
X
Y
Z
0.00
0.00 17.84
0.00
0.00
2.05

(m)

(n)

An example is depicted in Figure 5.4 for a three-span bridge with a prestressed concrete deck
point-supported on four columns. For the reasons of convenience mentioned in the previous
paragraph, the modal response spectrum analysis is normally done with the same nite
element model. The nodes of the deck are located at its mid-surface and normally have ve
DoFs each: three translations and two rotations (the rotation about an axis normal to the
mid-surface is normally not an independent DoF). The tributary mass of each node is
assigned to all its translational DoFs. Owing to the density of the nodes, no rotational mass
moment of inertia needs to be taken into account for the rotational DoFs.
Figure 5.4. Discretisation of a concrete slab deck in a rened mesh of shell nite elements

79

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

When the deck consists of a number of discrete parallel girders (whether of precast concrete or
steel, or a composite steelconcrete girder), a grillage-type of model is normally used for
gravity load analysis (including trafc loads). In such a model, intermediate nodes are introduced along each girder, and the girder is modelled as a spine of prismatic 3D beam/column
(sub)elements, having a T or I section with a ange width extending up to mid-distance to the
adjacent girder. Cross-(sub)elements, again of the prismatic 3D beam/column type, connect
each intermediate node along each spine to its counterparts on the spine(s) used for the
adjacent girder(s). They represent cross-beams wherever the deck has such elements, including
diaphragms over the supports on piers or abutments. In this case, the cross-(sub)element
has a T, I or inverted L section, similar to the real one, including any effective ange width
from the deck slab(s) at the top and sometimes at the bottom (see Figure 8.27 in Section
8.3.3.2 for an example of a deck consisting of two parallel composite girders with steel crossbeams). At all other intermediate nodes, any connecting cross-(sub)elements just represent the
transverse coupling of adjacent girder(s) by the deck slab at the top and at the bottom (if
there is one). Then, the cross-(sub)element has a rectangular section with an area and
moment of inertia about a horizontal axis equal to the sum of those of the top and bottom
deck slabs.
For concrete slab decks with parallel voids, a similar model may be used instead of a continuous
2D mesh of shell nite elements (in this case, with orthotropic properties reecting the effect of
the voids).
The grillage model of the deck highlighted above for gravity-load analysis may be used for modal
response spectrum analysis as well. To this end, tributary nodal masses are assigned to all three
translational DoFs of each node, but rotational DoFs have no rotational mass moment of
inertia, exactly as in the slab deck model with shell-type nite elements outlined two paragraphs
above. In this model, the extensional and exural rigidity of the deck as a whole was realistically
captured by the nite elements. In a grillage model it is not captured, unless special provisions are
made. The main concern is the in-plane stiffness of the deck. The extensional one is sufciently
represented, if each of the families of (sub)elements used in the longitudinal and the transverse
directions preserve on aggregate the cross-sectional area of the deck in these two directions.
The distribution across the width of the axial rigidities, EA, of the family of longitudinal
(sub)elements can approximate the in-plane exural rigidity of the deck, provided that an
effectively innite in-plane exural and shear rigidity is assigned to the cross-(sub)elements of
the transverse direction, so that these cross-(sub)elements remain normal to the longitudinal
axis of the deck (as in a bre model of a section, where the plane-sections hypothesis
converts

an in-plane distribution of bre axial rigidities, E dA, into a exural rigidity, Ey2 dA, of the
section). Finally, the in-plane shear rigidity of the deck in the transverse direction, GA, can be
effectively reproduced if the girder (sub)elements are assigned innite shear rigidity (zero shear
area) but their length (spacing of intermediate nodes along the girder) DL is about equal to
p
DL b [2(1 n)], where b is the spacing of the girders across the deck and n is the Poisson
ratio. In this way, the in-plane exural stiffness of a girder (sub)element (with its double xity
to the in-plane innitely stiff cross-beams), which equals 12E(tb3/12)/DL3, becomes equal to
G(tb)/DL.
Clauses 2.3.6.1(1)
2.3.6.1(3), 2.3.2.2(4)
[2]

Reecting the intended and expected elastic behaviour of concrete decks under the design
seismic action, Eurocode 8 species using in the linear analysis the properties of their full
uncracked gross section for the exural, shear and axial stiffness, except as prescribed in
Section 5.5.1.4 for continuity slabs over the joint between simply supported prefabricated
girders of adjacent spans. The same properties may also be used for reinforced concrete piers
in bridges designed for limited ductile behaviour. For those of bridges designed for ductile
behaviour, the effective exural stiffness is the secant-to-yield point (see Section 5.8.1).
Special provisions apply for the torsional rigidity of concrete decks, as this is reduced by
diagonal cracking (see Section 5.8.3).
5.5.1.3 Modelling of the connections between the deck and the piers
What has been said in Section 5.5.1.2 applies equally well when:
1

80

the deck is monolithically connected with the piers

Chapter 5. Modelling and analysis of bridges for seismic design

the deck is horizontally xed to the top of the piers through rigid bearings, but is free to
rotate there
3 movable bearings are placed between the top of the pier and the deck.
2

In case 1, the top of the pier and the deck immediately above it may share the same node, located
at the theoretical centre of the physical joint between the deck and the pier (or the pier column).
The part of the deck or the pier between that node and the faces of the physical joint (the end
sections of the deck or the pier) are modelled as rigid. Note that, unlike US codes (Caltrans,
2006), Eurocode 8 includes in the model the full width of the deck at and near monolithic
supports on pier columns as working in bending. In case 2, separate nodes are normally introduced: one at the top of the pier and another on the deck centroidal axis (or its mid-surface, if
the deck is a concrete slab as in the previous paragraph). These two nodes are constrained to
have the same translations but different nodal rotations. In case 3, separate nodes are introduced
at the top of the pier and at the deck immediately above it and are connected through a special
element modelling the bearing. This may be a truss element with extensional stiffness (kN/m) in
each direction in which the bearing presents exibility. Special bearings for seismic isolation have
nonlinear force-deformation behaviour, which should be modelled as such in the proper analysis
framework (i.e. with a nonlinear method). Although common laminated elastomeric bearings
also play a role as isolation devices, they are considered as elastic, and have a place in modal
response spectrum analysis. Their elastic stiffness is equal to EvAb/tq in the vertical direction
and to GAb/tq in the horizontal direction, where Ev is the effective elastic modulus of the elastomer at right angles to the lamination, G its shear modulus, Ab the horizontal surface area of
the bearing and tq the total thickness of the layers of elastomer in the bearing. The shear
modulus G can be considered as a material constant, with its value selected as outlined in
Section 6.10.4.1 of this Guide. By contrast, the effective elastic modulus Ev at right angles to
the lamination depends very much on its dimensions (its shape factor S and bulk modulus K,
etc, see Sections 6.10.4.1 and 6.10.4.2). The commonly used elastomeric bearings have relatively
high stiffness at right angles to the lamination, and may normally be assumed to be rigid in that
direction, at least to a rst approximation. If the bearing is xed (rigid) in one horizontal direction, the pier and deck nodes it connects are constrained to have the same displacement in that
direction.
Note that a deck monolithically or rigidly connected to certain piers or the abutment(s) may have
an out-of-phase response with respect to the piers or abutment(s) supporting it through movable
bearings. Such a response will be reected in higher modes; the increase in bearing deformations
and the pier forces it entails will be appropriately captured by the mode combination rules in
Section 5.5.4.
Shear keys are often provided on either side of the deck over a pier or abutment where the
deck is supported through movable bearings (e.g. sliding or elastomeric ones), in order to
prevent it from lateral unseating or fall-off. Closure of the gap between the deck and the shear
key under the design seismic action is a nonlinear phenomenon and cannot be realistically
reected in linear analysis: the secant stiffness allowed for seismic links in Part 2 of Eurocode
8, equal to the lateral resistance of a shear key divided by the travel of the deck until this
lateral resistance is attained, is hard to determine before dimensioning the shear key. To
bypass the problem, the gap may be lled by vertical elastomeric bearings placed between the
side of the deck and the shear key. A transverse stiffness of the spring equal to the aggregate
elastic stiffness, SEvAb/tq , of the vertical bearings of one side, may then be used at that pier or
abutment. If there is no gap or slack, this connection is equivalent to a xed bearing, and the
support transverse stiffness is that of the pier or abutment. A typical wall-type abutment has
quite high transverse stiffness and resistance, even discounting the uncertain lateral contribution of the soil/backll. Such a transverse connection of the deck to the abutment may be
taken as a rigid support.
Modelling and analysis of link slabs between deck spans of prefabricated
girders
A prefabricated deck (be it of a number of parallel girders or of a single box girder) is most often
simply supported on the piers. Elastomeric bearings are often used at both supports of the span,
although a xed bearing at one support and a movable one at the other have been used as an

Clause 6.6.1(2) [2]

5.5.1.4

Clause 4.1.3(3) [2]

81

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

option. This option, however, is not recommended, as it gives rise to several problems regarding
the seismic isolation of the deck, especially when used in conjunction with link slabs.
As pointed out in Section 4.2.1 of this Guide, a link slab (or continuity slab) is most often cast in
situ over a joint between the girders of adjacent spans (see Figure 4.1 for an example of the
detail). The link slab provides continuity to the top slab, to avoid a roadway joint and to
enhance motorist comfort, and creates a rigid diaphragm link between the adjacent deck
spans, so that the deck responds to the transverse seismic action as a single extended elastic
body over its entire length. Additionally, it serves as a sacricial seismic link between the two
spans against unseating and span drop-off. The link slab above the joint should be included in
the model. Its stretch between the end of a deck span consisting of simply supported girders
and that of the adjacent span may be modelled as a 3D beam element connecting the end
nodes of the beam elements modelling the deck (the ones above the support nodes at the pier
top). It is preferable to use for the link slab a single beam element along the bridge axis,
having a rectangular section with the vertical dimension the same as the true depth of the link
slab and the horizontal dimension as the full width of the deck slab. This applies even when
the girders of the deck span are not lumped into a single beam element along the bridge axis
but modelled individually, as in the grillage deck model highlighted in the penultimate paragraph
of Section 5.5.1.2. In this case, the single link slab element will connect the mid-point nodes along
the spine of beam elements modelling the diaphragm beams over the support nodes of each deck
span. The model should account for the vertical eccentricity between the centroidal axis of the
girders and the mid-plane of the link slab. To reect the serious damage and other implications
of the unavoidable concentration of inelastic deformations in the short stretch of the shallow link
slab above the joint, Part 2 of Eurocode 8 species for it an effective stiffness of 25% of
the uncracked concrete stiffness. The reduction factor of 0.25 should be incorporated both
in the out-of-plane exural rigidity, (EI )s , of the link slab element (for its exural behaviour
in the longitudinal direction of the bridge) and in its in-plane exural and extensional ones
(for the transverse response).
The following actions and action effects should be considered for the link-slabs:
1

For gravity loads and for imposed deformations:


The local bending due to wheel trafc loads on the slab, considered to be supported
beyond its clear span Lc , on the top slab of the rest of the deck (which is, in turn,
transversely supported on the main prefrabicated girders, see Figure 4.1).
The effects of the relative rotation, wd , of the end sections of the deck girders, one either
side of the link slab (Figure 5.5(a)). It induces a bending moment Mc,d (approximately)
constant over the length Lc , equal to
Mc;d

EI s wd
Lc

D5:7

where (EI )s is the out-of-plane exural rigidity of the link slab. Owing to redistribution
due to creep, the rotation wd is due not only to the permanent loads applied after
hardening of the link slab (due to the weight of the pavement, the sidewalks, etc.) but
also to a major part of the loads (about 75%) that were already acting before (the
weight of the deck girders and of the in-situ slab, if cast before the link slab).
A secondary effect is a tensile axial force of
Nc;d

GAb hd
w
tq d 2

D5:8

where G, Ab and tq are the shear modulus, the total horizontal area of the elastomeric
bearings under that span end and the total elastomer thickness of each bearing,
respectively, and hd is the depth of the deck girder (see Figure 5.5(a)). This axial force
should be added to the tensile axial forces that develop as a reaction of the elastomeric
bearings to the shortening of the deck due to concrete shrinkage and temperature
variation. Note that all of these effects of the gravity actions do arise from the analysis
82

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.5. Action effects in a link slab between precast girders due to imposed rotations: (a) from the
deck under gravity actions; (b) from the pier under seismic action
Lc
Mc,d
Beam axis

d /2

Beam axis

hd
Vb

Vb

Elastomeric bearing

Pier axis
(a)
Lc
Mc,p
Beam axis
Mc,p

hd

Beam axis

Elastomeric bearing

d = pLc

p
Pier axis
(b)

for them, if the link slab is included in the model as recommended above. If, in contrast,
it is excluded, they should be determined via Eqs (D5.7) and (5.8) from the value of wd
resulting from the analysis of the girders as simply supported under gravity actions.
2 For seismic loads:
For seismic action in the longitudinal direction, the main effect is induced by the
rotation, wp , of the pier head. As the link slab length, Lc , is much shorter than the spans
of the deck on either side, the latter may be approximately taken to remain horizontal.
So, the pier head rotation induces an equal chord rotation in the link slab; that is, skewsymmetric moments at its two ends (see Figure 5.5(b)) equal to
Mc;p

6EI s wp
Lc

D5:9

Skew-symmetric end moments equal to Mc,p are applied in turn to the girder deck. If Lg
is the girder length, the girder shear of 2Mc,p/Lg accompanying these moments, plus the
shear of 2Mc,p/Lc in the link slab, produce a vertical reaction in the bearings. If the two
rows of bearings under the girders of adjacent spans are at a centreline distance of Lb
along the deck axis, the couple of these reactions is translated into a total top moment
in the pier equal to

Mp 2Mc;p Lb

1
1

Lg L c


 2Mc;p

Lb
Lc

D5:10

This moment, acting on the pier top with a sign opposite to that of the cantilever
moment at the pier bottom, suggests that the link slab induces a secondary framing
83

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

effect between the pier top and the deck. Note that all of these seismic action effects
arise from the seismic analysis if the model includes the continuity slab as recommended
above and separate nodes at the top of the pier for the two rows of bearings at a
distance Lb apart. If it excludes these, they may be estimated via Eqs (D5.9) and (5.10)
from the value of wp resulting from the seismic analysis with the pier considered as a
vertical cantilever supporting the mass of the girders.
A secondary effect is an axial force due to the different stiffnesses of the piers and the
abutments in the longitudinal direction, which affects the (otherwise equal) horizontal
reactions of the elastomeric bearings to the inertia forces. This effect is minor, owing to
the large shear exibility of these bearings.
For seismic action in the transverse direction, the effects are usually minor: they are due
to the differences in stiffness of the individual piers and abutments, whose importance is
reduced by the dominant exibility of the bearings. However, higher mode effects are
possible in this case, due to the distribution of mass and transverse stiffness of the deck
over the full length of the deck. Nevertheless, the resulting seismic action effects usually
are not a problem for the link slab, owing to the large available section depth for the
deck in-plane bending (the full width of the deck). The seismic action effects in the link
slab deserve attention only if the bearings on either side of it (elastomeric, sliding or
other) have a very different stiffness in the transverse direction.

Clause 4.1.2(5),
Annex F [2]

5.5.1.5 Modelling hydrodynamic effects on immersed piers


5.5.1.5.1 Introduction
When bridge piers are immersed in water of considerable depth, hydrodynamic effects on the
seismic response to the horizontal seismic action are important. They are negligible as far as
the seismic response to the vertical component is concerned. For vertical piers, the effect for
the horizontal components may be estimated by means of an added mass of entrained water.
This approach, highlighted in the present section, is based on the following assumptions,
which are generally met when the seismic response of immersed piers is concerned:
g
g

water is incompressible
surface wave effects due to the seismic motion are negligible.

Hydrodynamic effects during the seismic response of an immersed structure are due to variations
in the water pressures on the waterstructure interface, and are, therefore, a function of the
motion of the structure relative to the water. The resultant forces of these pressures may be conveniently expressed as the product of an added mass matrix, Ma , considered as attached to the
structure, multiplied by the minus acceleration vector of the structure relative to the water in the
considered direction, U sw (i.e. MaU sw). The equation of motion of the structure is
MUA CU_ KU MaU sw

(D5.11)

where M, C and K are the mass, damping and stiffness matrices of the structure, respectively; for
hollow piers the mass of water enclosed within the pier is included in M; UA and U are the
displacement vectors of the nodes of the structure, absolute and relative to the ground, respectively; and U_ and U are the pseudo-velocity and pseudo-acceleration vectors of the structural
nodes, respectively.
The absolute, U A , relative, U and ground ug accelerations are related as U A U ug I, where the
unit vector I has all entries in the direction of the ground motion equal to 1, and 0 in all other
directions. So, the equations of motion become
MU CU_ KU MIug  MaU sw

(D5.12)

5.5.1.5.2 Added mass for the horizontal seismic components


Having zero shear stiffness, the water cannot follow a horizontal ground motion. Therefore, the
acceleration of the structure relative to the water is equal to its absolute acceleration: U sw U A .
Therefore, Eq. (D5.12) becomes
(M Ma)U CU_ KU (M Ma)Iu g
84

(D5.13)

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.6. Effect of the vertical component of water pressure on an immersed pier with non-vertical
outer surfaces: (a) destabilising on pier tapering downwards, M zPH bPV/2; (b) stabilising on pier
tapering upwards, M zPH bPV/2

PH /2

PH /2
PV /2

PV /2
z

PV /2

PH /2

PV /2
b

PH /2

This is equivalent to increasing the mass matrix to M Ma . As it lengthens the periods, the
added mass may reduce spectral accelerations; however, it increases the forces and displacements
of the seismic response. Its magnitude may be quite considerable, especially in hollow piers. A
pier of elliptical section and axes with length 2ax and 2ay subjected to a horizontal seismic
action component at an angle u to the x axis has an added mass per unit of immersed length
equal to
ma rp(a2y cos2u a2x sin2u)

(D5.14)

where r is the water density. Piers of circular cross-section have ax ay R. Informative


Annex F in Part 2 of Eurocode 8 gives the added mass of immersed rectangular piers in the
directions of its sides as a function of the aspect ratio of the pier section. Note that the added
mass of elliptical or rectangular piers is different in the two orthogonal principal directions of
the section.
The approach above gives the force resultant in the horizontal direction of the seismic
component and the related moments due to the variation in the water pressures on the lateral
faces of the pier. If those faces are not vertical, the horizontal component of the forces
resulting from these water pressures is well approximated by the simplied added mass
approach, but the vertical components of the pressures and their effects are not captured.
These vertical components produce a moment in the vertical plane of the horizontal seismic
component, which is destabilising (i.e. it increases the overturning seismic moment at the
base) if the outer faces of the pier taper downwards as in Figure 5.6(a); if they taper upwards
as in Figure 5.6(b), they are stabilising. If the non-vertical lateral surface of the pier is
symmetric with respect to a vertical plane at right angles to the horizontal seismic component, the vertical pressure components have a zero vertical force resultant. For immersed
piers with inclined lateral faces, it is recommended to use special nite element approaches
capable of dealing with the general coupled problem of the seismic response of an immersed
structure.
5.5.1.5.3 Vertical seismic component
Being incompressible, the water moves vertically with the ground. So, the acceleration of the
The added mass
structure relative to the water is equal to the one relative to the ground, U.
matrix associated with the vertical relative movement is denoted as Mav . An additional effect
of the water pressure variation on the lateral surface of the immersed pier is not due to the
motion of the structure relative to the water, and therefore not associated with an added
mass. It is due to the vertical acceleration of the water, ugv, which causes its apparent unit
weight to vary. The ensuing variation of the buoyancy force on the structure is conveniently
expressed as MbIugv , where Mb is the mass of the water displaced by the pier.
Before writing the equations of motion, the following should be pointed out:
g

Obviously the added mass of a pier in the vertical direction, Mav , has nothing to do with
that in the horizontal; indeed, if the lateral faces of the pier are vertical, Mav is zero.
85

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

A signicant added mass in the vertical direction derives from a boundary of the immersed
body in contact with water with a substantial projection on a horizontal plane (e.g. a
submerged pipe or tunnel supported on the piers).
Even for piers with inclined lateral faces (where Mav is non-zero), the relative acceleration
is due to the axial deformation of the pier; its magnitude is therefore
vector, U sw U,
much smaller than that of U sw corresponding to the horizontal directions where bending
deformation is involved.
The hydrodynamic effect due to the added mass in the vertical direction is very small for
piers of common bridges.

With the effects of added and buoyant masses included, Eq. (D5.12) is written as
(M Mav)U CU_ KU (M  Mb)Iu gv

(D5.15)

which shows that only the part M  Mb of the total vibrating mass M Mav is excited by ugv .
Therefore, for the analysis the mass matrix is increased to M Mav , but the modal excitation
factors must be modied to correspond to the excited mass M  Mb instead of the total.
Additionally, the sum of effective modal masses in the vertical direction is less than the total
structural mass.
In conclusion, it is justied for typical bridges to neglect the hydrodynamic effects of the vertical
seismic component, as allowed in Part 2 of Eurocode 8. In exceptional cases where this effect
should be included, either Eq. (D5.15) is solved, or recourse is made to special nite element
software.

Clauses 4.1.4(1),
4.1.4(3), 6.7.3(2) [2]
Clauses 4.2.3(1),
4.2.3(3), 5.4.2(1),
5.4.2(3), 6(1), 6(2),
Annex D [3]

5.5.1.6 Linear modelling of the foundation and the soil


Bridge foundations are either massive shallow footings, deep seated (caissons) or piled. For these
types of foundations, Part 5 of Eurocode 8 (CEN, 2004b) requires that soilstructure interaction
be taken into account in the seismic analysis. Bridge foundations are specically quoted as being
cases where considering soilstructure interaction is mandatory.
A rigorous treatment of linear soilstructure interaction is based on elasto-dynamic theories. A
direct (or complete) interaction analysis, in which both the soil and the structure are modelled
with nite elements is very demanding computationally and not well suited for design, especially
in 3D. A substructuring approach reduces the problem to more amenable stages, and does not
necessarily require repeating the entire analysis should the superstructure be modied. This
approach is of great mathematical convenience and rigour, stemming, in linear systems, from
the superposition theorem (Kausel and Roesset, 1974), according to which the seismic
response of the complete system may be computed in two stages (Figure 5.7):
Determination of the kinematic interaction motion, involving the response to base
acceleration of a system that differs from the actual one in that the mass of the
superstructure is equal to zero.
2 Calculation of the inertial interaction effects, referring to the response of the complete
soilstructure system to forces associated with base accelerations equal to the accelerations
arising from the kinematic interaction.
1

The second step is further divided into two subtasks:


g

computation of the dynamic impedances at the foundation level; the dynamic impedance
of a foundation represents the reaction forces acting under the foundation when it is
directly loaded by harmonic forces
analysis of the dynamic response of the superstructure supported on the dynamic
impedances and subjected to the kinematic motion (also called the effective foundation
input motion).

For rigid foundations, such as those of bridges, the dynamic impedances can be viewed as sets of
frequency-dependent springs and dashpots lumped at the underside of the footing or of the pile
cap. For rigid foundations, the complex-valued impedance matrix in the most general situation is
86

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.7. Superposition theorem in soilstructure interaction

Equivalent

Impedance function

Kinematic interaction

K=

(t)

M
V

Kxx

Kx
Kx K

(t)

a full 6  6 matrix. However, for regular geometries with two axes of symmetry and horizontally
layered soil proles, some off-diagonal terms are equal to zero, and the impedance matrix takes
the simplied form
2
6
6
6
6
6
K 6
6
6
6
4

k11

k15

0
0

k22
0

0
k33

k24

k42

k51
0

0
0

0
0

k44
0

0
k55

7
0 7
7
0 7
7
7
0 7
7
0 7
5
k66

D5:16

DoFs 1 to 3 refer to translations, and DoFs 4 and 5 to rocking about orthogonal horizontal axes
(Figure 5.8). DoF 3 is the vertical translation. Diagonal terms k33 , referring to this DoF, and k66
to the torsional DoF, are uncoupled with respect to the other DoFs. The two translational DoFs
are always coupled with the two rocking ones for piled or embedded foundations; this coupling
may be neglected only for surface foundations or when the embedment of the footing is shallow.
It may not be possible to introduce coupling (i.e. off-diagonal) terms to the stiffness matrix in
most commercial software codes. It is, however, feasible to account for coupling terms by translating the stiffness matrix to a point located at a depth z (.0) below the foundation where the
stiffness matrix becomes diagonal, and modelled with simple springs in all directions. For
instance, referring to directions 1 and 5, a diagonal stiffness matrix with a translational spring
stiffness of k11 and a rotational one of k55  z2k11 connected to the foundation at a depth
z k15/k11 via a rigid beam element produces the correct stiffness matrix at the foundation
(Figure 5.8).
Figure 5.8. Modelling of rotational-translational coupling in piles (a); DoFs and pile stiffness (b);
modelling of pile with uncoupled translational and rotational springs
3
Pile

55 = 1

11 = 1
6

K51

K55
K33

5
1

K11

Pile

K15

I=

z = K15/K11
K11

Rotational: K55 z2K11


(a)

(b)

87

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Analytical expressions for the different terms of the impedance matrix, Eq. (D5.16), for shallow
footings of various geometries, embedded foundations or pile foundations can be found in the
technical literature (Novak, 1974; Gazetas, 1983, 1991). Informative Annex C of Part 5 of
Eurocode 8 provides expressions for the stiffness of the head of individual piles for various
cases of dependence of the soil modulus on depth. Note that the pile head stiffness may be
strongly affected by pile group effects, reecting pilesoilpile interaction. This effect depends
on the cross-sectional size and spacing of piles, the soil properties and the frequency of excitation.
Unlike in static loading, it may be signicant even at large pile spacing. Simplied methodologies
have been developed to compute the group efciency under seismic loading (Dobry and Gazetas,
1988). For common pile sizes and spacing, the group effect on the stiffness typically varies from
2.5 to 6.
Each term in Eq. (D5.16) is a complex number that can be written as




vB d
s
d
d
s
d

k
k

i
a
c
k

i
c
kij
0 ij
ij
ij
ij
ij
vs ij

D5:17

where ksij is the static component of the impedance; the terms in parentheses, kdij and cdij,
correspond to the dynamic contribution to the impedance, and are frequency dependent;
a0 vB/vs is a dimensionless frequency; v is the circular frequency of the excitation; B is a
characteristic dimension of the foundation (radius, width, etc.); vs is the soil shear wave
velocity; and i2 1.
The terms in Eq. (D5.17) have the following physical interpretation: ksij kdij represents a spring and
ksij cdijB/vs a dashpot. The equivalent damping ratio of the system is

jij

d
vB c ij
vs 2k dij

D5:18

It is common and recommended practice in the context of modal response spectrum analysis
to bound upwards the value of the equivalent damping ratio from Eq. (D5.18) to 30%.
Although the substructure approach above is rigorous for the treatment of linear soilstructure
interaction, its practical implementation is subject to several simplications:
g

88

Fully linear behaviour of the system is assumed. However, it is well recognised that this is
a strong assumption, since nonlinearities are present in the soil itself and at the soil
foundation interface (sliding, uplift, gaping near pile heads, etc.). Soil nonlinearities may
be partly accounted for, as recommended in Part 5 of Eurocode 8, by choosing for the
calculation of the impedance matrix reduced values of the soil properties that reect the
soil nonlinear behaviour in the free eld (see Section 3.3.2.2 of this Guide). This implicitly
assumes that additional nonlinearities taking place at the soilfoundation interface (pile
shaft, edges of footing) do not affect signicantly the overall seismic response. For piles it
is recommended to further reduce the soil moduli at the top, starting from an almost zero
value at the pile head and increasing it to the full reduced value at a depth of 46 times
the pile mean cross-sectional dimension. Recommended reduction factors for the soil
moduli are given in Table 3.5 in Section 3.3.2.2.
Kinematic interaction is usually ignored. This means that the input motion used for the
dynamic response of the structure is simply the free-eld motion. Although that
assumption is exact for shallow foundations subject to vertically propagating body waves,
and partly exact for shallow foundations in a more complex seismic environment or for
exible piles, it becomes very inaccurate for embedded or very stiff piled foundations. In
particular, the true input motion to embedded foundations consists not only of a
translational component but of a rotational one as well.
The frequency-dependent terms of the stiffness matrix are approximated as constant.
Except in the very rare case of a homogeneous soil prole, this condition is far from being
fullled. A fair approximation for the constant value of the stiffness term is one
corresponding to the frequency of the soilstructure interaction mode. To nd such a

Chapter 5. Modelling and analysis of bridges for seismic design

Figure 5.9. Example of a simple rheological model for foundation impedance

value, iterations on the spring constant should be carried out for each vibration mode,
until the chosen value corresponds to the computed frequency. More rened models,
involving additional masses and springs, are also possible to represent the frequency
dependence of the impedance term. One such example is shown in Figure 5.9. The model
parameters may be tted to the exact variation of the impedance term (as exemplied in
Figure 5.10), or assessed from simplied assumptions based on cone models (Wolf and
Meek, 2004). Alternatively, frequency domain solutions may be used: they offer several
advantages, such as a rigorous treatment of the frequency-dependent impedance matrix
and full account of the radiation damping represented by the imaginary part of the
impedance matrix. Special software tools have been developed to solve soilstructure
interaction problems in the frequency domain (Lysmer et al., 2000), widely known in the
industry.
A wide class of so-called Winkler models, based on the concept of springs and dashpots to
model the effect of the soil on the foundation have been used extensively. They represent the
Figure 5.10. Example of the accuracy of the rheological model of Figure 5.9 used to model the exact
dynamic rocking impedance of the Rion Antirrion bridge foundation (Pecker, 2006)
1.E+08
Model
Finite element analysis
Stiffness: MN/m

5.E+07

0.E+00

5.E+07

1.E+08
0

1.E+07

2
3
Frequency: Hz

Model
Finite element analysis

Dashpot: MN s/m

8.E+06
6.E+06
4.E+06
2.E+06
0.E+00
0

2
3
Frequency: Hz

89

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 5.11. Modelling techniques for linear soilstructure interaction: (a) impedance matrix; (b) Winklertype model

Superstructure

6 1 kinematic
motion time
series
Condensed 6 6
stiffness matrix [K]

Pile cap

Condensed 6 6
mass matrix [M]

(a)

Horizontal
motion 1
Horizontal
motion 2

Superstructure

Depth varying
free field
motions

Pile cap
kh1
kh2

Horizontal
motion 3

kh3

Horizontal
motion 4

kh4

Horizontal
motion n

khn

kv1

Vertical
motion 1

kv2

Vertical
motion 2

kv3

Vertical
motion 3

kv4

Vertical
motion 4

kvn

Vertical
motion n

Pile
foundation

(b)

interaction with the soil using springs and dashpots distributed across the footing or along the
pile shaft. Although conceptually the soil reaction forces are still represented by the action of
springs and dashpots, unlike the impedance matrix approach, there is no rational or scientically
sound method for the denition of these springs and dashpots. Their values, but more importantly their distribution over the foundation, vary with frequency; there is no unique distribution
reproducing the global foundation stiffness for all degrees of freedom. For instance, although
highly questionable, a uniform distribution may be chosen for the vertical stiffness, but cannot
accurately match the rocking stiffness. Furthermore, two additional difculties arise for pile
foundations:
g
g

the choice of the springs and dashpots should reect the pile group effect
as the seismic motion varies with depth, different input motions should be dened at all
nodes shared between the piles and the soil; one should resort to a separate analysis for
the determination of these input motions.

Therefore, in view of all the uncertainties underlying the choice of their parameters, Winkler-type
models, although attractive, should not be favoured. Figure 5.11 depicts both types of modelling:
g
g

that based on the concept of the impedance matrix, which is rigorous in the framework of
linear systems
the approximate Winkler-type model.

5.5.2
Modal analysis results
In modal response spectrum analysis, the modal shapes (eigenmodes) in 3D and the natural
frequencies (eigenvalues) are rst computed. In a full 3D model, each mode shape in general
has displacement and rotations in all three directions, X, Y and Z, at all nodes i of the model.
The eigenmodeeigenvalue analysis gives for each normal mode, n:
1
2
3

The natural period, Tn , and the corresponding circular frequency, vn 2p/Tn .


The mode shape vector Fn .
Factors of the participation of the mode to the response to the seismic action component
in direction X, Y or Z, denoted as GXn , GYn or GZn , respectively. GXn is computed as

GXn FTn MIX/FTn MFn SiwXi,nmXi/Si(w2Xi,nmXi w2Yi,nmYi w2Zi,nmZi)

(D5.19)

where i stands for nodes associated with DoFs, M is the mass matrix, IX is a vector with
elements of 1 for the translational DoFs parallel to direction X and all other elements
90

Next Page
Chapter 5. Modelling and analysis of bridges for seismic design

equal to 0, wXi,n is the element of Fn corresponding to the translational DoF of node i


parallel to X and mXi is the associated element of the mass matrix. Similarly for wYi,n ,
wZi,n , mYi and mZi . If M contains rotational mass moments of inertia, IuXi,n , IuYi,n and
IuZi,n , the associated terms are included in the sum at the denominator of GXn . The
denitions of GYn , GZn are similar.
4 The effective modal masses in directions X, Y and Z are MXn , MYn , and MZn , respectively:
MXn (FTn MIX)2/FTn MFn (SiwXi,nmXi)2/Si(w2Xi,nmXi w2Yi,nmYi w2Zi,nmZi); similarly for
MYn and MZn. They give the peak force resultant in mode n along direction X, Y or Z:
VbX,n Sa(Tn)MXn , VbY,n Sa(Tn)MYn or VbZ,n Sa(Tn)MZn , respectively. The sum of
effective modal masses in X, Y or Z over all modes is equal to the total mass.
Peak modal seismic action effects in the response to the seismic action component in direction X,
Y or Z may be calculated as follows:
For each mode n the spectral displacement, SdX(Tn), is calculated from the design
acceleration spectrum of the seismic action component; for example, for direction X as
SdX(Tn) (Tn/2p)2SaX(Tn).
2 The nodal displacement vector of the structure in mode n due to the seismic action
component of interest, let us say direction X, UXn , is computed as UXn= SdX(Tn)GXnFn .
3 Peak modal values of the effects of the seismic action component of interest are computed
from the modal displacement vector of step 2; member modal deformations (e.g.
curvatures or chord rotations) directly from the nodal displacement vector of mode n;
modal internal forces at member ends by multiplying the member modal deformations by
the member stiffness matrix; etc.
1

The so-computed peak modal responses are exact but occur at different instances in the response,
and can be combined only approximately. Section 5.5.4 presents statistical combination rules for
the peak modal responses.
5.5.3
Minimum number of modes
Modal response spectrum analysis should take into account all modes contributing signicantly
to any response quantity of interest. This is hard to achieve in practice. As the number of modes
to be considered should be specied as input to the eigenvalue analysis, either a very large number
may have to be specied from the outset or the eigenvalue analysis may have to be repeated with a
higher number of modes requested this time.

Clauses 4.2.1.2(1),
4.2.1.2(2) [2]

Eurocode 8 considers the total force resultant in the direction of each seismic action component
as the prime response quantity of interest, and sets as a goal of the eigenvalue analysis the capture
of at least 90% of its full value. This is translated into a criterion for the participating modal
mass: Eurocode 8 requires the N modes that are taken into account to provide together a total
participating modal mass along any individual direction of the seismic action components considered in the design, at least 90% of the total mass. In the example of Figure 5.3, three modes
sufce for the longitudinal direction, but nine are needed for the transverse; this requires
computing up to the 49th eigenmode of the bridge.
If the above criterion is hard to meet, Part 2 of Eurocode 8 allows as an alternative the computation of all modes with periods above 0.033 s, provided that they collectively account for at
least 70% of the total mass in the direction of each component of interest. Then, we should
make up for the missing mass by amplifying all computed modal seismic action effects by the
ratio of the total mass to that accounted for so far. If the computed modes with periods
above 0.033 s do not collectively capture at least 70% of the total mass, more modes should
be computed, until the minimum of 70% is achieved. The amplication of computed modal
results follows. Clearly, this option is so conservative and uneconomic for the design that the
designer is motivated to go after as many modes as necessary to capture at least 90% of the
total mass.
5.5.4
Combination of modal results
In modal response spectrum analysis, it is convenient to take the elastic response in two different
modes as independent of each other. The magnitude of the correlation between modes i and j

Clause 4.2.1.2(3) [2]

Clause 4.2.1.3 [2]

91

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.119

Chapter 6

Verication and detailing of bridge


components for earthquake resistance
6.1.

Introduction

The rules for the verication of the strength and global ductility of bridge components in Section 5
of Part 2 of Eurocode 8 (CEN, 2005a), alongside those for detailing the components for local
ductility in Section 6, elaborate and implement the compliance criteria for the no-collapse
requirement set out in Section 2 and in Chapter 2 of this Designers Guide. These rules are
different for bridges with ductile or with limited ductile behaviour. Different rules apply for
bridges equipped with seismic isolation: they are given in Section 7 of Part 2 of Eurocode 8,
and elaborated and exemplied in Chapter 7 of this Designers Guide.
For better illustration and comparison, the rules for the two design alternatives of ductile or
limited ductile behaviour are presented here in tabular form (Table 6.1). Some general aspects
and concepts are highlighted rst, and the rationale behind certain verication procedures and
rules is presented.

6.2.

Combination of gravity and other actions with the design


seismic action

Both at the local level (for the verication of members and sections) and at the global level (for
Clauses 3.2.4(1) [1]
the calculation of masses), the design seismic action is combined with other actions as specied in Clauses 5.5(1), 5.5(4),
4.1.2(3) [2]
EN 1990:2002 (Basis of design CEN, 2002) for the seismic design situation and elaborated for
bridges in Part 2 of Eurocode 8. Symbolically this combination is
Ed Gk Pk AEd c2,1Q1,k Q2

(D6.1)

where Gk is the characteristic (nominal) value of the permanent actions (normally the self weight
and all other dead loads), Pk is the characteristic value of prestress after losses, AEd is the design
seismic action (i.e. that for the reference return period times the importance factor of the
bridge), Q1,k is the nominal value of trafc loads, c2,1Q1,k is the combination value of trafc
loads (normally the quasi-permanent, i.e. arbitrary-point-in-time, trafc loads) and Q2 is the
quasi-permanent value of variable actions with long duration (earth or water pressure,
buoyancy, currents, etc.).
Coefcient c2,1 is dened in normative Annex A2 of EN 1990:2002 as a Nationally Determined
Parameter (NDP). As already pointed out in Section 5.4 of this Guide in connection with
Eqs (D5.6), Part 2 of Eurocode 8 recommends the following values:
g
g
g

c2,1 0 for bridges with normal trafc and footbridges


c2,1 0.2 for road bridges with severe trafc conditions (dened in a note as motorways
and other roads of national importance)
c2,1 0.3 for railway bridges with severe trafc conditions (dened in a note as intercity
rail links and high-speed railways).

Part 2 of Eurocode 8 recommends taking only the characteristic value, qk , of the uniform trafc
load (UDL) of Load Model 1 (LM1) with all its nationally applying adjustments as contributing to the quasi-permanent part of severe trafc loads. Note that, at least as far as the
calculation of masses is concerned, the small values of c2,1 even for bridges with heavy trafc
119

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 6.1. Eurocode 8 rules for dimensioning and detailing of bridge piers and deck

q factor

Limited ductile

Ductile

q  1.5 (see Table 5.1)

q  1.5 (see Table 5.1)

Eurocode 2 Class B or C
Eurocode 2 Class B or C

Eurocode 2 Class C

Limitations for reinforcing bars


In plastic hinges
Outside plastic hinges

Ultimate limit state (ULS) strength verication and dimensioning of the deck
ULS in exure

MGd M Ad(1)  M yd(2)

MGd aCDM Ad(1),(3)  M yd(2)

ULS in shear

VGd qV Ad(1)  VRd /g Bd1(4)

VGd aCDV Ad(1),(3)  VRd /gBd(5)

ULS strength verication and dimensioning of the piers in exure


o
(i) In plastic hinges
MGd MAd DM 2nd-order(1),(6)  M Rd(7)
(ii) Outside plastic hinges

MGd MAd DM 2nd-order(1),(6)  M Rd(7)

ULS strength verication and dimensioning of the piers in shear


o
(i) In plastic hinges
VGd qV Ad(1)  VRd /gBd1(4)
(ii) Outside plastic hinges

VGd aCDV Ad(1),(3)  VRd,o /gBd(9)

MC  M Rd(7),(8)

VGd aCDV Ad(1),(3)  VRd /gBd(5)

Detailing of pier columns for ductility within length Lh from end section of plastic hinge
Length Lh from the end section where full connement and antibuckling reinforcement is required:
o
(i) If hk NEd/Acfck  0.3
Lh max[hc; Ls/5](10),(11),(12),(13)
Lh 0.3LsMRd /(1.3MEd)(10),(11)
(ii) If 0.3 , hk  0.6
Lh 1.5max[hc; Ls/5](10),(11),(12),(13)
Transverse reinforcement(14) for connement:(15),(16)
Maximum hoop/tie spacing along pier

min(hc; bc)/5 or Dc/5(13)

Rectangular hoops and cross-ties:


mechanical ratio in each direction
Asw/(shbc)(fyd/fcd)(17)

.0.28hkAc/Acc 0.13(rL  0.01)fyd/f cd(19)


.0.08

.0.37hkAc/Acc 0.13(rL  0.01)fyd/f cd(19)


.0.12

Circular hoops or spirals: volumetric


mechanical ratio, 4Asp/(shDsp)(fyd/fcd)(18)

0.39hkAc/Acc 0.18(rL  0.01)fyd/f cd(19)


.0.12

0.52hkAc/Acc 0.18(rL  0.01)fyd/f cd(19)


.0.18

Hoop legs or cross-ties engaging bars:


spacing along straight sides of section

200 mm
min(hc; bc)/3(13)

Connement reinforcement at one-half


its amount from 0 to Lh

Between Lh and 2Lh from end section

Transverse reinforcement(14) for restraint of vertical bars against buckling:


Maximum hoop/tie spacing along pier
Area of restraining tie-legs per linear
metre of pier column length

ddbL, where d 2.5ftk/fyk 2.25, with d  5, d  6(20)


P
AsLfyL/(1.6fyw)(21)

Vertical bars (index L) throughout the height of pier columns(22)


Minimum steel ratio, rmin(23)

0.1Nd/Acfyd, 0.2%

Maximum steel ratio, rmax

4%

Diameter, dbL 

8 mm

Distance of unrestrained bar from


nearest restrained along perimeter

150 mm

Transverse bars (index w) over the part of pier columns where special detailing for ductility is not required (22)
Diameter, dbw 

dbL/4

Spacing, sw 

400 mm, 20dbL, min(h; b) or D

At lap splices, if dbL . 14 mm: sw 

240 mm, 12dbL, 0.6 min(h; b) or 0.6D

120

Chapter 6. Verication and detailing of bridge components for earthquake resistance

around the clock are consistent with the action of vehicles on the deck as tuned mass dampers
that reduce the global response of the bridge, as has been shown in shaking table tests.
In the special but common case where only elastomeric bearings resist the seismic action, the
action effects of imposed deformations (thermal actions, shrinkage, support settlement,
ground movement due to seismic faulting, etc.) should also be included in the combination of
Eq. (D6.1). Note, however, that these effects are only relevant for the design of the bearings
themselves, which is normally based on the deformations not the forces imposed on the
bearings by actions. Additionally, in this case the behaviour under the design seismic action is
linear elastic; so is the analysis with q 1.0. Therefore, the action effects of imposed deformations do not vanish, as normally taken in design for the ultimate limit state (ULS) after the
presumed redistribution of self-equilibrating action effects.

Clauses 5.5(2),
5.5(3) [2]

(1) MAd, VAd: moment and shear from linear analysis for the design seismic action, Ad. MGd, VGd: moment and shear from linear analysis for the
quasi-permanent gravity loads concurrent with Ad
(2) Myd: design yield moment of the deck section at yielding of the tension chord, taking into account any prestressing. For bending about the
vertical axis, Myd may be taken to occur when the longitudinal reinforcement yields at a distance from the edge of the deck section of 10%
its depth for bending in a horizontal plane (i.e. 10% of the width of the top slab) but not beyond the outer web of the section
(3) aCD: capacity design magnication factor. It may be approximated as the ratio of the sum of pier seismic shears when overstrength moment
resistances develop at all potential plastic hinges in each pier to the sum of pier shears from the linear analysis for the design seismic action
(Eq. (D6.9)). The seismic shear at overstrength moment resistance at the plastic hinges of a pier is computed assuming the simultaneous
presence of the effects of the quasi-permanent loads concurrent with the design seismic action
(4) gBd1: Nationally Determined Parameter safety factor with the recommended value gBd1 1.25. VRd: design shear resistance per Eurocode 2,
taking into account any prestressing
(5) It is allowed to take gBd equal to the value of gBd1 in note (4) above or to subtract from it qVEd/VC,o  1 (but not to a nal value below
gBd 1), where VEd is the maximum shear from the analysis for the seismic design situation and VC,o is the capacity design shear without the
upper limit VC,o  qVEd
(6) DM2nd-order is the additional moment due to second-order effects. It may be taken equal to (1 q)dEdNEd/2, where NEd is the axial force and
dEd is the relative horizontal displacement of the two pier ends in the seismic design situation (see Eq. (D6.3))
(7) MRd: design moment resistance of the pier section, for the concurrent value of the pier axial force and the orthogonal moment component
from the analysis for the seismic design situation
(8) MC: linear bending moment diagram between the overstrength moments at the ends of the pier column where plastic hinges may form (see
Figure 6.1)
(9) VRd,o: design shear resistance per Eurocode 2 computed for a truss inclination of 458 and using as dimensions of the section those of the
conned concrete core to the centreline of the perimeter hoop
(10) Ls M/V: shear span at the end section according to the analysis for the seismic design situation
(11) The maximum length for the two directions of the pier section applies
(12) MEd: moment at the pier end section from the analysis for the seismic design situation. MRd: as in note (7) above, for the pier end section
(13) hc, bc, Dc: depth, width and diameter of the conned core to the centreline of the perimeter hoop
(14) Anchored with 1358-hooks. When hk  0.3, a 908 hook with a 10-diameter extension is allowed at one end of a cross-tie, at alternating
ends in adjacent cross-ties horizontally and vertically. Long cross-ties with 1358 hooks at both ends may be lap-spliced between the ends
(15) Connement is not required in those pier columns where:
g hk NEd /Acfck  0.08 or
g a curvature ductility factor value of
13 for ductile behaviour or
7 for limited ductile
can be achieved in the plastic hinge without connement. This is possible in piers with wide compression anges, e.g. in hollow rectangular
piers, in the weak direction of wall-like piers, etc. On this basis, connement is not required in hollow rectangular piers with
hk NEd/Acfck  0.2
(16) If connement all along the long dimension of a wall-like pier is not necessary per the second bullet point in (15) above, but is required in
the pier strong direction, it should be provided up to a distance from the edges where the strain drops below 50% of the ultimate strain of
unconned concrete: 1cu2 0.0035
(17) Asw/sh: total cross-sectional area of hoop and tie legs at right angles to the dimension bc of the concrete core (measured to the outside of
the perimeter hoop) per unit length of the pier column
(18) Asp: cross-sectional area of the spiral or hoop bar. sh: pitch of the spiral or spacing of the hoops. Dsp: diameter of the spiral or hoop bar
(19) Ac: area of gross concrete section. Acc: area of conned core to the perimeter hoop centreline. hk NEd/Acfck. rL: ratio of vertical
reinforcement
(20) dbL: vertical bar diameter. ftk/fyk: ratio of the characteristic tensile strength to the characteristic yield stress of vertical bars. Eurocode 2 limits
the lower 10% fractile of ft/fy to (ft/fy)k  1.15 for steel Class C and (ft/fy)k  1.08 for steel Class B; these values may be used in lieu of ftk/fyk,
in the absence of steel-specic information
P
AsL: total cross-sectional area of vertical bars restrained
(21) fyw: yield stress of tie. fyL: yield stress of vertical bars.
p at . 908 hooks or corners of
the ties (tie-legs restraining the bar at a 458 corner contribute too with their cross-sectional area divided by 2).
(22) These rules are per Eurocode 2 and apply both to ductile and limited ductile design
(23) Strictly, the minimum steel ratio in pier columns is that of Eurocode 2, because Part 2 of Eurocode 8 does not specify a minimum steel ratio
for pier columns. According to Part 1 of Eurocode 8, the minimum steel ratio of 1% applies only to concrete columns in buildings, and is
certainly high for pier columns. Such a high minimum ratio may be the source of irregular seismic behaviour if the piers have the same
section but very different seismic demands (see Sections 5.10.2.1 and 5.10.2.2). Nonetheless, it is often applied in practice by designers who
consider the minimum steel ratio specied in Eurocode 2 for any type of concrete column as too low for earthquake resistance. Note that
the fundamental requirement of Part 2 of Eurocode 8 for a minimum of local ductility in all ductile elements is of essence. In this respect, the
amount of steel reinforcement should be sufcient to provide a design moment resistance of the pier section, MRd, not less than the cracking
moment: Mcr Wc(fctm NEd/Ac), where Wc is the elastic section modulus, fctm is the mean tensile strength of concrete, NEd is the axial
force from the analysis for the seismic design situation (positive if compressive) and Ac is the area of the concrete section. Normally, a steel
ratio of 0.5% provides this minimum of local ductility, without causing signicant irregularity

121

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

6.3.

Verication procedure in design for ductility using linear


analysis

In force-based seismic design with linear analysis employing a q factor greater than 1.0, the
verications outlined in the following sections are carried out.
6.3.1

Verication of the exural resistance of plastic hinges at the ULS

Clauses 5.4(1),
Flexural plastic hinges are dimensioned so that their design moment resistance, MRd , exceeds the
5.6.2(1), 5.6.3.1(1) [2] moment from the analysis for the seismic design situation, MEd , including second-order effects, if
Clauses 5.2.4(1)
applicable:
5.2.4(3), 6.1.3(1),
(D6.2)
MEd  MRd
7.1.3(1) [1]

Linear analysis is carried out for the combination of the actions of Eq. (D6.1). Superposition
does apply, so the seismic action effects from an analysis for the design seismic action alone
may be superimposed on those from the analysis for the other actions in Eq. (D6.1). Approximations may be used for the second-order effects in the seismic design situation. Part 2 of
Eurocode 8 invites the National Annex to introduce sufciently accurate methods to include
them. Such methods should account for the less ominous character of second-order effects of
seismic displacements compared with those due to persistent displacements, owing to their
reversed nature. In a non-binding note, Part 2 of Eurocode 8 considers it sufcient to neglect
these effects in linear analysis and simply to increase the moment calculated from the analysis
at pier ends where plastic hinges are expected to form by the following additional moment
(Paulay and Priestley, 1992):
DM2nd-order (1 q)dEeNEd/2

(D6.3a)

where NEd is the axial force and dEe is the horizontal displacement of the top relative to the base
of the pier from the analysis for the design seismic action. According to Paulay and Priestley
(1992), an increase in the moment resistance by the amount of Eq. (D6.3a) compensates for
the absorbed energy loss due to the second-order effects of the peak seismic displacements,
mddEe (with md taken equal to q, see Eq. (2.1a)). Note that, by mistake, Part 2 of Eurocode 8
uses in Eq. (D6.3a) the total design displacement in the seismic design situation, dEd , instead
of the correct dEe; dEd is the sum of (a) the displacements dE , due to the design seismic action
alone (see Eq. (D5.48)), (b) dG , due to the quasi-permanent actions on the bridge, and
(c) c2dT , due to the thermal actions present in the seismic design situation. Equation (D6.3a)
accounts for second-order effects due to the seismic displacements; to account for those due to
displacement components (b) and (c), Eq. (D6.3a) may be generalised as follows:
DM2nd-order [(1 q)dEe/2 dG c2dT]NEd

(D6.3b)

In tall and slender piers, dG should include:


g
g
g

a possible displacement (eccentricity) d0 due to the permanent loads


the displacement (eccentricity) dimp due to geometric imperfections, according to clause 5.2
in EN 1992-2 (CEN, 2005b)
the effect of concrete creep (through the creep coefcient w).

A convenient approximation that captures all three effects above on the basis of the nominal
stiffness method per clause 5.8 of EN 1992-1-1 (CEN, 2004b) is


1w
D6:4
dG d0 dimp 1
n1
where
n NB/NEd

(D6.5a)

with
NB p2(EI )eff/L2o
122

(D6.5b)

Chapter 6. Verication and detailing of bridge components for earthquake resistance

denoting the buckling load estimated according to the nominal stiffness method of clause 5.8 of
EN 1992-1-1, using the effective stiffness (EI )eff of the pier section in the seismic design situation
(see Section 5.8.1), and Lo is the effective length of the pier column in the direction considered.
Guidance for the value of Lo is given in Section 4.4.3 of this Guide.
The value of MRd in Eq. (D6.2) should be calculated according to the relevant rules of the
pertinent material Eurocode. It should be based on the design values of material strengths;
that is, the characteristic values, fk , divided by the partial factor gM of the material. Being key
safety elements, the partial factors, gM , are Nationally Determined Parameters, with values
dened in the National Annexes to Eurocode 8. Eurocode 8 itself does not recommend the
values of gM to be used in the seismic design situation: it just notes the options of choosing
either the values gM 1 appropriate for the accidental design situations, or the same values as
for the persistent and transient design situation. This latter option is very convenient, because
the plastic hinge may then be dimensioned for the largest design value of the action effect due
to the persistent and transient or the seismic design situation. With gM 1, the plastic hinge
will have to be dimensioned once for the action effect due to the persistent and transient
design situation and then for that due to the seismic design situation, each time using different
values of gM for the resistance side of Eq. (D6.2).
In piers, Eq. (D6.2) is checked using as MEd the algebraically maximum and minimum value of
each uniaxial moment component from the analysis and as MRd the design resistance for the
concurrent value of axial force, NEd , and the acting moment in the orthogonal direction of the
pier in the seismic design situation. If a separate analysis is carried out for each horizontal
component of the seismic action, X and Y, and the outcomes are combined via Eqs (D5.2),
the verication is also done separately for the moment components from Eq. (D5.2a) and for
those from Eq. (D5.2b). By contrast, Eq. (D5.1) gives only the peak absolute values of the
two moment components for the seismic design situation. These values are not concurrent,
and it is too conservative to take them as such. Statistical approaches have been proposed in
Gupta and Singh (1977) and elaborated and extended in Fardis (2009) for the estimation
of the axial force and orthogonal moment components likely to take place concurrently with
the peak value of each moment component under the simultaneous action of the two or
three independent components of the seismic action.
6.3.2
Capacity design of regions, components or mechanisms for elastic response
Regions outside the exural plastic hinges, and non-ductile mechanisms of force transfer within
or outside the plastic hinges, as well as components for which absolute protection from damage
(e.g. the deck, normally the foundation, seismic links, etc.) is desired under the design seismic
action are dimensioned to remain elastic until and after all potential exural plastic hinges
form in the bridge. This is pursued by overdesigning:
(a) all regions outside exural plastic hinges
(b) the non-ductile structural components or mechanisms of force transfer
(c) all parts that are meant to remain elastic (including the deck and the foundation)

Clause 5.6.1(1) [2]


Clause 4.3.3.5.1(2)c
[1]

Clauses 2.3.4(1)
2.3.4(3), 2.3.6.2(2),
5.3(1), 5.6.2(2)a,
5.6.3.6(1), 5.7.2(1),
5.8.2(3), 6.5.2(1),
6.5.2(2), 6.6.2.1(1),
6.6.3.1(3) [2]

relative to the corresponding action effects, Ed , from the linear analysis for the seismic design
situation. Non-ductile structural components in the context of (b) include seismic links, xed
bearings, sockets and anchorages for cables and stays, etc.; for concrete, the non-ductile
mechanism of force transfer is by shear. In bridges of limited ductile behaviour, the overdesign
is limited to case (b) and to the foundation; it is normally accomplished by multiplying the seismic
action effects from the linear analysis by q and carrying out the ULS strength verications with
the increased action effects. In bridges of ductile behaviour, by contrast, capacity design is
employed to prevent the components listed above under (a) to (c) from becoming inelastic. In
capacity design, all potential plastic hinges are presumed to develop overstrength moments,
goMRd. Then, simple analysis is employed to estimate the action effects that develop in the
regions to be protected from inelasticity when this complete plastic mechanism forms.
Normally, equilibrium sufces for this.
6.3.3
Detailing of plastic hinges for ductility
Flexural plastic hinges are detailed to provide the deformation and ductility capacity that is consistent with the demands placed on them by the design of the structure for the chosen q factor.
123

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

6.4.
Clauses 5.3(3) 5.3(4),
5.3(6) [2]

Capacity design of regions other than exural plastic hinges in


bridges of ductile behaviour

6.4.1
Overstrength moments of exural plastic hinges
In general, capacity design effects are established separately for seismic action in each one of the
longitudinal and the transverse directions of the bridge, each time in the positive or negative
sense. The design moment resistance, MRd , entering in the overstrength moment,
Mo goMRd , refers to the end section of the plastic hinge, and is calculated assuming that the
concurrent values of axial force, NEd , and of the moment in the orthogonal direction of the
pier column (as these are obtained from the linear analysis for the seismic design situation and
for the considered direction longitudinal or transverse and sense of the seismic action
positive or negative) are acting together with MRd . The overstrength factor, go , is meant to
take into account the uncertainty in material strengths and the hardening of the section
between yielding and ultimate strength. It is a Nationally Determined Parameter, with the
following recommended values:
g
g

for steel members:


go 1.25
for concrete members:

go 1.35
go 1.35[1 2(hk  0.1) ]
2

if hk NEd/(Ac fck)  0.1

(D6.6a)

if hk NEd/(Ac fck) . 0.1

(D6.6b)

with NEd from the linear analysis for the seismic design situation and the considered
direction and sense of the seismic action.
6.4.2

Estimation of capacity design effects in the plastic mechanism from the


overstrength moments of the exural plastic hinges
Clauses 5.3(1), 5.3(6), Normative Annex G of Part 2 of Eurocode 8 outlines a multistep procedure for the estimation of
G.1(1)G.1(5) [2]
the capacity design effects assuming the complete plastic mechanism forms. First, the overstrength moments, Mo goMRd , of the exural plastic hinges are established. Next is the
calculation at each plastic hinge of the difference between the overstrength moment and the
moment, MG , produced at the end section of the plastic hinge by the non-seismic actions in
the seismic design situation:
DMh goMRd  MG

(D6.7)

Bending moments enter this calculation with the signs they have in the seismic design situation
for the considered direction and sense of the seismic action. If, by contrast, goMRd and DMh
are always taken as positive, MG is also positive if it acts on the section in the same sense
as goMRd does. The physical sense of the action of goMRd and DMh should be the same as
that of the rotation of the plastic hinge: for example, at the two ends of a pier column that
is xed against rotation at both the top and bottom, they are such that the pier column is in
counterexure.
For the general case, Part 2 of Eurocode 8 requires the capacity design effects to be determined
separately for each horizontal direction of the seismic action, and each sense of the inertial forces
in each direction (positive or negative). This is to take into account the effect on the moment
resistance of the plastic hinge sections of (a) the magnitude of the axial force and (b) any
asymmetry in the shape or reinforcement of the section (a rarity in pier columns). The sense of
the seismic action component materially affects the magnitude of the axial force if the pier has
more than one column in the direction of this component (e.g. in twin-blade piers see
Section 4.2.2.2 of this Guide under the longitudinal seismic action).
The next step is the estimation of the change in action effects in the plastic mechanism, DAC ,
when the moments of exural plastic hinges increase from MG to goMRd (i.e. by DMh). Often,
this can be done on the basis of equilibrium alone. In the simple example of Figure 6.1 (a pier
column in counterexure with overstrength moments at its two ends), DAC amounts to an
increase from a linear moment diagram up the column to another linear diagram and from
one constant value of shear force to another. In the nal step, the action effect increments,
124

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Figure 6.1. Capacity design moments in a pier column forming plastic hinges at the top (index t) and
bottom (index b). Dashed line: moments ME from the analysis for the seismic design situation. Solid line:
capacity design moments, MC, with a cut-off near each end at the design moment resistance, MRd, as
controlled by the reinforcement of the end section and the axial load from the analysis for the seismic
design situation
M E,t

MRd,t

Mo,t = 0M Rd,t

Deck
Plastic hinge

MC

Pier

Plastic hinge

Lh
Mo,b = 0M Rd,b

M Rd,b

M E,b

DAC , are superimposed on the effects of the non-seismic actions in the seismic design situation,
AG . The outcome of the superposition is the capacity design effects:
AC AG DAC  AG qAE

(D6.8)

where AE denotes the generic effect of the design seismic action from the linear analysis.
Section 8.2.9 of this Guide exemplies the application of the general procedure above to estimate
the capacity design effects in a bridge deck monolithically connected to the piers, for longitudinal
seismic action.
Normally, simplications of the general procedure are employed. Simplication is indeed
encouraged by normative Annex G of Part 2 of Eurocode 8, as long as it respects equilibrium.
A very common simplication entails using equilibrium to estimate the seismic shear
accompanying the increase in the base moment of pier column i from MbG,i to goMbRd,i (by
DMbh,i) and of its top moment from MtG,i to goMtRd,i (by DMth,i) where the index t is used for
the top, and b for the bottom. If the top is supporting the deck through a hinge, or if we are
dealing with the transverse direction, then DMth,i 0. Capacity design gives for the seismic
shear of pier column i
DVC;i

b
t
DMh;i
DMh;i
 qVE;i
Hi

Clauses G.2(1),
G.2(2) [2]

D6:9

which is superimposed on the shear due to gravity and other permanent actions in the seismic
design situation, VG , to give the total capacity design shear in pier column i:
VC,i VG DVC,i

(D6.10)

VE,i in Eq. (D6.9) is the shear of pier column i from the linear analysis under the design seismic
action alone.
Another simplication is to neglect the moments and shears in pier column i due to gravity and
other permanent actions: MG 0, VG 0. With this usually good approximation, the
capacity design shears are the same, no matter the sense of the seismic action and this distinction between positive or negative is no longer necessary.
By far the most important simplication is to assume that the capacity design effects are
proportional to the total shear force in all pier columns of the bridge, despite the departure
125

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

from linearity that marks the development of the plastic mechanism with the gradual formation
of exural plastic hinges. This assumption facilitates greatly the estimation of the capacity design
effects in the deck, in any seismic links and on foundations and abutments, etc., as
P
V
DAC P C;i AE  qAE
VE;i

D6:11

where the summations are over all pier columns, AE and VE.i denote the generic effect and the
shear in pier column i, respectively, due to the design seismic action from the linear analysis,
and DAC is the generic capacity design effect, to be superimposed on the generic effect of the
non-seismic actions in the seismic design situation, AG , according to Eq. (D6.8).
The above, as well as the overall framework for the capacity design of bridges, conforms with the
position of Eurocode 8 (both Parts 1 and 2) that capacity design takes place separately in the
longitudinal and transverse directions. In case, however, a single modal response spectrum
analysis is carried out for both simultaneous horizontal seismic action components and (the
more rigorous and convenient) Eq. (D5.1) is employed to estimate the peak values of the
seismic action effect, one cannot associate AE and VE,i in Eq. (D6.11) to a single component:
they encompass the effects of both. In that case, the ratio SVC,i/SVE,i may be computed separately from the shears in the two directions of the bridge; the minimum of the two ratios is used
then in Eq. (D6.11) to amplify all the seismic action effects, AE . Indeed, it is the minimum of the
two ratios that is associated with the earliest development of a full plastic mechanism in the
bridge.

Clauses 5.6.3.1(2),
5.6.3.2(1) [2]

6.4.3
Capacity design moments outside the exural plastic hinges of ductile piers
As depicted schematically in Figure 6.1, the vertical reinforcement of ductile pier columns stays
the same all along the nominal length, Lh , of the plastic hinge. In case the section of the pier
column decreases from the pier base upwards, the moment resistance of the pier column,
MRd , may also decrease a little as we go up within the nominal length of the plastic hinge (an
additional reduction may be due to a reduction in the axial force, NEd , from the pier base
upwards). Outside that length, Part 2 of Eurocode 8 allows the termination of some vertical
bars according to the linear diagram of capacity design moments, MC , which connects the overstrength moments, goMbRd,i and goMtRd,i at the two end sections of the pier column. This diagram
provides a signicant margin for any increase in the pier moments due to higher modes that
reect the distributed mass of the piers, even when the analysis ignores this mass and such
modes. Note, however, that, in practice, part of the vertical reinforcement is terminated
between the two ends of only fairly tall piers.
6.4.4

Clauses 5.6.3.5.1(1),
5.6.3.5.2(1),
5.6.3.5.2(2) [2]

Capacity design shear in a joint between a ductile pier column and the deck
or a foundation element next to a column plastic hinge
High shear stresses in the core of the joint between a pier column and its foundation element
(spread footing or pile cap) accompany the transfer of the bending moment at the base of the
column to the foundation. High shear develops also in the joint region between the deck and
the top of the column, if the pier column is monolithically connected there to the deck. The
joint may fail under these stresses, and its failure will be brittle. So, if the bridge is designed
for ductile behaviour, the shear stresses in such joints are determined via capacity design, for
the worst possible condition: when a plastic hinge has formed at the end section of the pier
column and has developed its overstrength moment, Mo goMRd , established according to
Section 6.4.1.
In the following, the pier column is indexed by c. The horizontal component monolithically
connected to it whether a deck or a foundation element is called, for simplicity, beam
and indexed by b.
Let us denote by xz the vertical plane within which bending of the pier column takes place (see
Figure 6.2). A nominal shear stress is calculated as the average within the joint core; it has the
same value no matter whether it is parallel to the horizontal axis x and acts on a horizontal
plane normal to the vertical axis z (tzx , denoted for simplicity vx) or is parallel to the vertical
axis z and acts on a vertical plane normal to the horizontal axis x (txz , denoted for simplicity

126

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Figure 6.2. Forces on a joint between a pier column (c) and the deck (b) above a column plastic
hinge. (Turn the gure upside down for a joint with a foundation element.)
Vjz

z
y
x

Njz
Vb2c

Plastic hinge

Njx

Vb1c

zb

hb

0FtRc
0MRdc
zc
hc

vz). vx is obtained by smearing a joint horizontal shear force Vjx over the horizontal crosssectional area of the joint, while vz results from smearing a vertical shear force Vjz over its
vertical cross-sectional area bjzb , normal to the plane of bending of the column:
vx

Vjx
bj z c

vz

Vjz
bj z b

vx vz ; vj

D6:12

If the pier column has a depth hc in the plane of bending and width bc orthogonal to it (for a
circular column of diameter Dc , conventionally bc hc 0.9Dc), the effective width of the
joint at right angles to the plane of bending of the pier column is
bj bc 0.5hc

b j  bw

(D6.13)

where bw is the physically available width of the deck or the foundation element (the beam)
parallel to bc . The depth dimensions of the horizontal or vertical cross-sectional area of the
joint in Eq. (D6.12) are the internal lever arms of the column and the beam end sections, zc
and zb , respectively (see Figure 6.2). For convenience, zc and zb may be taken as 90% of
corresponding effective section depths:
zc  0.9dc

zb  0.9db

(D6.14)

The nominal shear stress vj of Eq. (D6.12) is normally determined from that of the shear forces,
Vjx , Vjz , which is in the direction of the yielding component. Unlike buildings, where the yielding
component is normally the beam(s), in bridges it is the column. Then, capacity design gives for
the design vertical shear of the joint, Vjz ,
Vjz goFtd,c  Vb1C  goMRd,c/zc  Vb1C

(D6.15)

where Ftd,c is the tension force resultant in the column section corresponding to the design
exural resistance MRd,c , go is the overstrength factor from Section 6.4.1 and Vb1C is the shear
force of the beam adjacent to the tensile face of the column (taken as positive if acting in the
sense shown in Figure 6.2). The design horizontal shear of the joint Vjx may then be calculated
from Eq. (D6.12) as Vjx Vjzzc/zb .
The value of Vb1C should correspond to the capacity design effects of the plastic hinge, according
to the general procedure in Section 6.4.2. Its magnitude is small compared with the rst term in
Eq. (D6.15), and is normally estimated on a case-by-case basis. For example:
1

In the longitudinal direction, at a monolithic connection of one out of nc columns of a pier


to the deck: Vb1C  goMRd,c/Ld  Vb1G/nc (normally negative), where Ld is the average
span length on either side of the pier and Vb1G is the total deck shear due to quasipermanent loads at the face of the pier.
127

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

At the connection of a spread footing with


a single-column pier in the transverse direction
any pier column in the longitudinal direction:
Vb1C  0, assuming that the weight of the uplifting part of the footing beyond the column
compensates for any soil pressures on its underside.
At the connection of a pile cap with
a single-column pier in the transverse direction
any pier column in the longitudinal direction:
Vb1C  goMRd,c/zpc  0.5NEd , with NEd denoting the axial force of the pier column in the
seismic design situation and
zpc (Bp/3)np/(np  1) if np is odd or
zpc (Bp/3)(np 1)/np if it is even,
where Bp is the distance between the extreme piles of the group in the vertical plane of
bending of the column and np is the number of piles in each row of piles parallel to that
vertical plane.
In the transverse direction, at a monolithic connection of a multi-column pier to the deck:
Vb1C  0, as the most adverse joint is that of the column with the largest value of MRd,c
(i.e. of the outer one where the seismic action induces a compressive axial force in the
pier); then, Vb1C arises from the width of the deck beyond the outer column, and is very
small.
In the transverse direction, at the connection of a multi-column pier to a spread footing:
the joint below every single column may have to be considered, from the one nearest the
uplifting edge of the footing, where the value of MRd,c is the lowest (because the seismic
action induces a tensile axial force) but Vb1C  0 with the same reasoning as in case 2
above, to the column closest to the down-going edge, where MRd,c is the largest (owing to
the compressive seismic axial force) but Vb1C may be large; the values of Vb1C next to that
column or any intermediate one may be determined from force equilibrium under the axial
forces in the columns in the seismic design situation, NEd , and the soil pressures on the
underside of the footing when the overstrength moments, goMRd,c , develop at the base of
the columns (the soil pressures taken with a linear distribution across the footing).
In the transverse direction, at the connection of a multi-column pier to a pile cap: the joint
below every column should be considered, from the column nearest to the uplifting edge,
where the tensile seismic axial force reduces the value of MRd,c but Vb1C is low or about
zero, to the opposite one where the compressive seismic axial force increases MRd,c but
Vb1C may be large; the values of Vb1C next to every column may be estimated from force
equilibrium under the axial forces in the columns in the seismic design situation, NEd , and
the pile reactions when the overstrength moments, goMRd,c , develop at the base of the
columns.

6.4.5

Clauses 5.3(7),
5.3(8) [2]

Capacity design effects in piers or abutments of ductile bridges supporting


the deck on sliding or elastomeric bearings
In a ductile bridge, where certain piers are rigidly connected to the deck (with monolithic
connection, xed bearings or seismic links without slack or clearance), some other piers or the
abutments may support the deck on sliding or elastomeric bearings. The shears and moments
in these piers or abutments are calculated assuming that the bearings develop the following
capacities:
g

For sliding bearings: a horizontal force of gofRdf , where Rdf is the maximum design
friction force, equal to the maximum friction coefcient multiplied by the maximum
vertical reaction in the seismic design situation and gof 1.30.
For elastomeric bearings: a horizontal force equal to gof 1.30 multiplied by the
horizontal stiffness of the bearing, GbAb/tq , multiplied by the maximum bearing
deformation corresponding to the total design displacement of the deck, dEd , at the
horizontal level where it is seated on the bearing. The value of dEd is estimated from
Eq. (D6.36), according to point 3 in Section 6.8.1.2.

Coefcient gof with a value of 1.30 is meant to account for potential hardening of the bearing due
to ageing, etc., from its installation (possibly as a replacement) until the seismic event.
128

Chapter 6. Verication and detailing of bridge components for earthquake resistance

6.5.

Overview of detailing and design rules for bridges with ductile


or limited ductile behaviour

As mentioned in Section 6.1, Table 6.1 lists the rules for the verication of strength and the
capacity design of the deck and the piers, as well as the rules for the detailing of plastic hinges
for ductility. In addition, the detailing rules per Eurocode 2, applying outside the plastic hinge
region, are given. Supplemented with a long list of footnotes, the table is almost a standalone
design aid.
Most of the detailing rules for Eurocode 8 listed in Table 6.1 are prescriptive and unlike the
important concept of capacity design are not given much attention in the text of the present
chapter.

6.6.

Verication and detailing of joints between ductile pier


columns and the deck or a foundation element

6.6.1
Stress conditions in the joint
The joint core is considered to be in a triaxial stress state, with shear stresses tzx txz ; vj given
by Eq. (D6.12)(D6.15) and normal stresses sx ; nx , sy ; ny , sz ; nz equal to
nx

Njx
bj hb

ny

Njy
hb hc

nz

Njz
bj hc

Clauses 5.6.3.5.2(3),
5.6.3.5.2(4) [2]

D6:16

Njx and Njy are normal forces in the plane of the horizontal element (deck or foundation element)
into which the pier column frames; they are within and at right angles, respectively, to the plane
of bending of the pier column (see Figure 6.2 for the convention of the axes). They derive from
any prestressing of the horizontal element that may be effective in the joint core (after losses),
plus any other in-plane normal forces in the seismic design situation (the average of analysis
results at opposite vertical faces of the column). The vertical normal force at the centre of the
joint is
Njz

bc
N
2bj Gc

D6:17

where NGc is the axial force in the column due to the quasi-permanent actions in the seismic
design situation; the factor 2 in the denominator reects the reduction in the normal stresses
due to NGc from NGc/(hcbc) at the end section of the column to zero at the opposite face of
the horizontal element. All stresses and forces in Eqs (D6.16) and (D6.17) are taken as
positive for compression.
6.6.2
Verication of the integrity of the joint
The major threat to the joint comes from crushing of its core by diagonal compression. Part 2 of
Eurocode 8 considers that the diagonal strut running through the joint core within the vertical
plane xz will fail in compression upon exceedance of the compressive strength of concrete, as
this is reduced due to tensile strains at right angles to the strut and may be enhanced by the
conning effect of any compressive stress ny normal to the xz plane and any closed reinforcement
transverse to this plane. In the classical variable strut inclination model of Eurocode 2 for
concrete members with shear reinforcement, diagonal compression failure should be checked
against a shear resistance normalised to the cross-sectional dimensions of the web of


VRd;max
fck MPa
0:5  0:6 1 
D6:18
fcd sin 2u
bw z
250

Clauses 5.6.3.5.3(2),
5.6.3.5.3(3) [2]

where u is the strut inclination, and the brackets and the factor 0.6 are for the reduction of the
uniaxial compressive strength of concrete due to tensile strains at right angles to the strut. The
maximum resistance is for u 458. So, Part 2 of Eurocode 8 adopts the following simple verication criterion against diagonal compression failure of the joint:


f MPa
vj  0:3 1  ck
ac fcd
250

D6:19
129

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

where vj is the average shear stress in the joint from Eqs (D6.12)(D6.15), while

ac 1 2(njy ry fsd)/fcd  1.5

(D6.20)

accounts conservatively for the increase in the compressive strength of the diagonal strut
resulting from any conning pressure (ny , from Eq. (D6.16)) and/or reinforcement (ry) in the
transverse direction y; ry Asy/(hchb) is the reinforcement ratio of any closed stirrups at right
angles to the vertical plane xz of the joint, taken to work with a reduced stress,
fsd 300 MPa, in order to limit the width of cracks parallel to the xz plane. The outcome of
Eq. (D6.19) is on the safe side compared with the upper bound of joint shear strength demonstrated by the available cyclic tests on unconned beamcolumn joints (see Fardis, 2009).
6.6.3

Dimensioning of the joint reinforcement

Clause 5.6.3.5.3(3) [2] For the purposes of calculating the joint reinforcement, Part 2 of Eurocode 8 adopts for the joint

core the variable strut inclination model of Eurocode 2, which also underlies Eqs (D6.18) and
(D6.19). It again uses the value u 458 for the strut inclination. This gives for the joint core
the following total amounts of vertical (z) and horizontal (x) reinforcement in the vertical
plane of column bending:
Asx (Vjx  Njx)/fyd ,

Asz (Vjz  Njz)/fyd

(D6.21a)

or, in terms of steel ratios, normalised to the same joint core areas as the corresponding forces,

rx ; Asx/(bjhb) (vj  nx)/fyd


6.6.4

rz ; Asz/(bjhc) (vj  nz)/fyd

(D6.21b)

Maximum and minimum reinforcement in the joint

Clause 5.6.3.5.3(4) [2] Yielding of the joint reinforcement is a ductile failure mode. By contrast, diagonal concrete

crushing, if vj exceeds the limit value of Eq. (D6.19), is brittle. If one places more reinforcement
than required by Eqs (D6.21), the shear stress, vj , that may develop in the joint core may increase
to the limit of Eq. (D6.19), inviting diagonal concrete crushing. For this reason, Part 2 of
Eurocode 8 controls the amount of horizontal joint reinforcement, so that vj from Eq. (D6.19)
is not reached, even when the benecial effects of transverse compression and connement are
discounted:

rx ; Asx =bj hb  rmax

ry ; Asy =hb hc  rmax



f MPa fcd
where rmax 0:3 1  ck
fyd
250
D6:22

Clauses 5.6.3.5.3(1),
5.6.3.5.3(6) [2]

The joint will rst crack when the principal tensile stress, sI , under the system of normal stresses,
sx ; nx , sy ; ny , sz ; nz from Eqs (D6.16) and shear stresses tzx txz ; vj from Eqs (D6.12)
(D6.15), exceeds the (design value of ) tensile strength, fctd . Setting sI equal to fctd gives the
joint shear stress at diagonal cracking of the joint according to Part 2 of Eurocode 8:
vj;cr

s



nx
nz
1
fctd
1
 1:5fctd
fctd
fctd

D6:23

where fctd fctk0,05/gc 0.7fctm/gc . Theoretically, the joint may rely on the tensile strength of
concrete if vj from Eqs (D6.12)(D6.15) is less than the joint cracking stress, vj,cr . However, in
ULS verications we normally do not rely on the tensile strength of concrete, placing instead
minimum reinforcement that can provide alone the tensile stresses released in case cracking
occurs for any reason. Part 2 of Eurocode 8 derives on the basis of Eqs (D6.21b) and (D6.23)
discounting the benecial effects of any transverse compression the minimum joint reinforcement ratio as

rx ; Asx/(bjhb)  rmin

ry ; Asy/(hbhc)  rmin

where

rmin fctd/fyd

(D6.24)

The minimum and maximum reinforcement ratios apply in both horizontal directions, even when
framing action with the column takes place only in one of them.
130

Chapter 6. Verication and detailing of bridge components for earthquake resistance

6.6.5
Detailing of the vertical reinforcement at joints
The vertical bars of the pier column should be fully anchored into the horizontal element,
extending over its full depth even when it is not necessary for their anchorage. In addition,
excepting those not serving as extreme tension reinforcement in any of the two horizontal
directions of bending, they should have a standard 908 hook inwards or a head at the end (see
Figure 6.3(a)).

Clauses 5.6.3.5.4(1),
5.6.3.5.4(2),
5.6.3.5.4(4)
5.6.3.5.4(6) [2]

The vertical reinforcement of the joint should be in the form of stirrups clamping inwards the
bars of the horizontal element at its outer (far) face. At least 50% of them should be fully
closed stirrups, engaging both the top and the bottom bars of the horizontal element; the rest
may be U bars, engaging the bars only at the outer (far) face, but not at the face where the
pier column frames into (see Figure 6.3(a)).
To reduce the congestion of reinforcement in the central part of the joint in plan, the density of
vertical stirrups may be reduced there, but not below the minimum ratio of Eqs (D6.24). The
balance, DAsz , may be placed outside the central area in plan but next to it. If framing action
of the column with the beam takes place in the vertical plane xz, DAsz may be placed within
the applicable joint width bj in the y direction and not beyond 0.5hb from each column face
along the x direction (Figure 6.3(b)). If there is framing action within the vertical plane yz as
well, this operation takes place separately in that direction, with the values of DAsz and bj now
applicable (Figure 6.3(c)). Vertical stirrup legs in the overlapping areas count for both directions
of framing action.

Figure 6.3. Alternative arrangement of joint reinforcement: (a) vertical section within the plane xz;
(b) joint plan view, if plastic hinges form only in the x direction; (c) joint plan view if plastic hinges form in
the x and y directions

#50%

Aszb

hb

Beam-column
interface

Asx

hb /2

hc

hb /2

(a)
Stirrups in common areas
count in both directions

hb /2

hb /2

bj

hb /2
x

Areas for
placing Aszb

(b)

hb /2

(c)

131

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Clauses 5.6.3.5.4(1),
5.6.3.5.4(3),
5.6.3.5.4(5),
5.6.3.5.4(7) [2]

6.6.6
Detailing of joint horizontal reinforcement
Up to 50% of the top and bottom bars in the horizontal element may count towards the required
horizontal joint reinforcement area, Asx , provided they are continuous through the joint or fully
anchored beyond it. The rest of Asx should consist of stirrups or hoops (preferably of the same
shape and diameter as those in the column plastic hinge area) enclosing the column vertical
reinforcement as well as the ends of beam horizontal bars anchored into the joint.
To reduce the congestion of reinforcement in the central part of the joint in elevation, the total
area of horizontal bars may be reduced there by DAsx  DAsz , but not below the minimum ratio
of Eqs (D6.24). The balance, DAsx, should then be placed at the top and bottom faces of the horizontal element over and beyond its reinforcement for the verication in exure under capacity
design effects. The additional bars should be placed within the joint width bj and be adequately
anchored in order to be fully effective at a distance hb from the column face (Figure 6.3(a)).

6.7.
Clause 4.2.4.4(2) [2]

Verications in the context of design for ductility based on


nonlinear analysis

6.7.1
Format of verications
If the analysis for the seismic design situation is nonlinear, regions of the bridge intended to have
elastic behaviour (notably the deck, parts of the piers outside the exural plastic hinges, shear
keys and other critical connections, etc.) are veried to remain in the elastic range. In principle,
such verication may take place in terms of either forces or deformations. It is consistent with the
verications for linear analysis, if it is in terms of forces. Then, exure with or without axial load
may be veried as for linear analysis, with material partial factors applied on the resistance side.
Components intended to stay elastic (including the foundation and seismic links) and brittle
modes of force transfer and behaviour in ductile components (notably shear in piers rigidly
connected to the deck), are also veried in terms of forces. Their design resistances are
computed as for linear analysis; that is, with partial factors and, in addition, divided by the
Nationally Determined Parameter safety factor, gBd1 , which divides the shear resistance in
bridges with limited ductile behaviour and has a recommended value gBd1 1.25 (see
Table 6.1). Alternatively, however, the value gBd 1.0 may be applied if capacity design
effects are used, computed with the minimum value of the overstrength factor recommended in
Part 2 of Eurocode 8 (go 1.35 or 1.25 for concrete or steel piers, respectively, see Section 6.4.1).
Flexural plastic hinges in the piers are veried in terms of deformations, namely plastic hinge
rotations, u pl. As pointed out in informative Annex E of Part 2 of Eurocode 8, the plastic part
of a hinge rotation at a pier end may be taken as equal to the plastic part of the chord
rotation at that end. Similar to all verications for nonlinear analysis (including those of the
two paragraphs above), the demand is the value from the nonlinear analysis for the seismic
design situation (in this case the value of the plastic hinge rotation computed, u pl
E ). The
?
question is, then, how much is the corresponding design capacity, u pl
Rd
6.7.2
Plastic hinge rotation capacity
pl
Part 2 of Eurocode 8 is not very specic regarding u pl
Rd . As a design value, u Rd should be obtained
pl
by dividing the expected ultimate value, u um , by a safety factor, gR,pl , that reects the natural
variability of materials and components, model uncertainty and/or experimental scatter. Part
2 of Eurocode 8 mentions as possible sources for u pl
um relevant test results or calculation from
ultimate curvatures. In a non-binding note it mentions also informative Annex E as a source
of information for u pl
um and gR,pl .

Annex E [2]

In the absence of specic justication based on actual data, informative Annex E is content with
the value gR,pl 1.40. However, gR,pl is meant to also reect model uncertainty; so its value
should depend on the model used for the calculation of u pl
um .
The model presented in informative Annex E of Part 2 of Eurocode 8 is the classic point hinge
model, based on curvatures, f, and on the plastic hinge length, Lpl:



Lpl
Ls 
uu fy fu  fy Lpl 1 
D6:25
3
2Ls

132

Chapter 6. Verication and detailing of bridge components for earthquake resistance

where uu is the ultimate chord rotation and Ls the shear span (moment-to-shear ratio) at the end
of the member. The rst term on the right-hand-side of Eq. (D6.25) is the elastic part of uu; the
second is the plastic part, upl
u.
Section analysis is the basis for the calculation of the yield curvature, fy , and of the ultimate
curvature, fu . Both should be determined for the value of the section axial load from the
nonlinear analysis under the seismic design situation. Informative Annex E of Part 2 of
Eurocode 8 refers to the construction of a full momentcurvature curve for this value of the
axial load until the value of fu is reached and to an elasticperfectly plastic approximation of
this curve that maintains the same plastic deformation energy (i.e. the same area under the
curve) from rst yielding of the reinforcement up to fu (see Figure 5.13(a)). The curvature at the
corner of this elasticperfectly plastic curve may be taken as fy . The value of fu is dened on
the basis of the failure criteria of the section, and is independent of the construction of the full
momentcurvature curve (in fact, the value of fu denes a priori the end point of the curve),
while the values of uu and upl
u are not very sensitive to the exact value of fy . Therefore, there is
little point in constructing the full momentcurvature curve a sensitive and often onerous undertaking just for the sake of the value of fy . This is underlined further by Part 2 of Eurocode 8,
which itself suggests (in informative Annex C) the values for fy given by Eqs (D5.57) and (5.58)
in Section 5.8.2. The same expressions may well be used for the purposes of Eq. (D6.25).
Informative Annex E of Part 2 of Eurocode 8 proposes using in the section analysis for the
calculation of fu an elasticlinearly strain-hardening stressstrain (s1) law for the reinforcing
steel: a mean yield stress higher by 15% than fyk; a mean tensile strength, ftm , higher by 20%
than the characteristic value, ftk; and the strain at ftm equal to the characteristic strain value at
maximum stress, 1uk . If the values of ftk and 1uk of the specic reinforcing steel used are
unknown, they may be taken as equal to the minimum values specied in Annex C of
Eurocode 2 for the steel class used.
As far as the s1 law of concrete is concerned, informative Annex E of Part 2 of Eurocode 8
proposes the one in Mander et al. (1988), with an ultimate strain of 0.35% for unconned
concrete. The ultimate strength of conned concrete in Mander et al. (1988) is adopted, with a
value according to Elwi and Murray (1979) of
fc fc 1 K

D6:26

where
K 2:254

r

p
2p
1 7:94  1 
fc
fc

D6:27

and is taken to occur at a strain of (Richart et al., 1928)


1co  1co(1 5K )

(D6.28)

where 1co 0.002 is the strain at the ultimate strength of unconned concrete, fc . The conning
pressure in Eq. (D6.27) is
D6:29

p 0:5arw fyw
where a is the connement effectiveness factor (Mander et al., 1988):
g

for a spiral with pitch s and diameter Dc ,


a1

s
2Dc

D6:30a

for circular hoops with diameter Dc and spacing s,




s
a 1
2Dc

2
D6:30b
133

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

for vertical bars laterally restrained at distances bi (i indexes the bars) along a rectangular
perimeter tie with sides bcx , bcy and spacing s,
P 2 !



bi =6
s
s
a 1
D6:30c
1
1
2bcx
2bcy
bcx bcy

fyw is the yield stress and rw is the volumetric ratio of conning reinforcement with respect to the
conned core to the centreline of the conning hoop, spiral or perimeter tie:
g

for a spiral or circular hoops of cross-sectional area Asp , pitch s and diameter Dc ,

rw
g

4Asp
sDc

D6:31a

for rectangular ties with total cross-sectional areas Aswx , Aswy at right angles to sides bcx,
bcy of the concrete core, respectively, and spacing s,
s
Asx Asy
rw 2
sbcx sbcy

D6:31b

Informative Annex E of Part 2 of Eurocode 8 adopts the ultimate strain of conned concrete in
Paulay and Priestley (1992):
1cu 0:004 2:81su;w

p
fc

D6:32

with p from Eqs (D6.29)(D6.31), f c from Eqs (D6.26) and (D6.27) and 1su,w the strain at the
tensile strength of the conning reinforcement.
For values of fy and fu established as above, informative Annex E of Part 2 of Eurocode 8
proposes the following empirical expression for Lpl , in terms of the shear span, Ls , the
characteristic yield stress of the vertical bars in MPa, fyk (not the mean, fym 1.15fyk , used in
the calculation of fy and fu) and their diameter, dbL:
Lpl 0.1Ls 0.015fykdbL

(D6.33)

If Ls/d , 3 (where d is the effective depth of the pier section), informative Annex E of Part 2 of
p
Eurocode 8 multiplies the second term of the right-hand-side of Eq. (D6.25) by (Ls/3d). The
outcome is the nal expected value of u pl
u , which is further divided by gR,pl 1.40, to yield a
design plastic rotation capacity, u pl
d.
Figure 6.4 compares the predictions of the above procedure (but without dividing u pl
u by
gR,pl 1.40) with the experimental values of the ultimate total chord rotation, uu , in a large
number of cyclic tests to exural failure. The comparison is shown separately for (a) compact rectangular sections (constituting the vast majority of the experimental data), (b) long rectangular
(wall-like) sections, (c) hollow rectangular or other anged sections and (d) circular piers. With
the exception of the very low (safe side) predictions for circular piers, the average agreement is
good (overall median experimental-to-predicted ratio of 1.045, including the circular piers),
but the scatter of test results about the predictions is high (the overall coefcient of variation
of the experimental-to-predicted ratio is 66.5%).
Figures 6.5(a) to 6.5(c) compare the same data of Figures 6.4(a) to 6.4(c) with the predictions of
the expressions given in Part 3 of Eurocode 8 (CEN, 2005c) and Biskinis and Fardis (2010) for
(a) beam/column rectangular sections, (b) rectangular walls sections and (c) non-rectangular
walls or other anged sections. Figure 6.5(d) compares the same data as those in Figure 6.4(d)
with the predictions of the expressions proposed in Biskinis and Fardis (2012) specically
for circular piers. There are small differences between the expressions used in each of
Figures 6.5(a) to 6.5(c); by contrast, the difference between the expression underlying
Figure 6.5(d) and those used for Figures 6.5(a) to 6.5(c) is fundamental. Owing to these
134

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Figure 6.4. Experimental cyclic ultimate chord rotation versus predictions of approach in informative
Annex E of Part 2 of Eurocode 8: (a) compact rectangular sections, (b) rectangular walls; (c) hollow or
anged rectangular members; (d) circular piers
25
Median:
u,exp = 0.99u,pred

15

10

Median:
u,exp = 1.35u,pred

u,exp: %

u,exp: %

20

10
15
u,pred: %

20

25

2
3
u,pred: %

(a)

(b)

16
14

Median:
u,exp = 2.53u,pred

12
u,exp: %

u,exp: %

10
8
6

4
Median:
u,exp = 1.21u,pred

4
u,pred: %
(c)

2
6

8 10
u,pred: %

12

14

16

(d)

differences, the agreement is more balanced and good for circular piers as well. The overall
median of the experimental-to-predicted ratio is 1.00. The scatter, although signicant, is
much less than in Figures 6.4(a) to 6.4(d) (the overall coefcient of variation of the experimental-to-predicted ratio is 36.5%).

6.8.

Overlap and clearance lengths at movable joints

6.8.1
Minimum overlap length
6.8.1.1 General
A deck supported on a pier or abutment via a horizontally movable device (slider, elastomeric
bearing or special isolation device) should be prevented from dropping off by means of a
minimum horizontal overlapping of the underdeck and the top of the supporting abutment or
pier in any direction along which relative displacement between the two is physically possible.
The same applies at a movement joint separating the deck between adjacent piers or a pier
and an abutment, where one part of the span is vertically supported on the other (normally
the shorter on the longer of the two).
6.8.1.2 Minimum overlap length at an abutment
According to Part 2 of Eurocode 8, the minimum overlapping length of the end of the deck
(Figure 6.6) seated on an abutment is
min lov  lm deg des

Clauses 2.3.6.2(1),
2.3.6.2(3), 6.6.4(1),
6.6.4(2) [2]

Clauses 6.6.4(1)
6.6.4(3), 2.3.6.3(2),
3.3(6) [2]

(D6.34)
135

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 6.5. Experimental cyclic ultimate chord rotation versus predictions of expressions in Biskinis and
Fardis (2012) for circular piers or in informative Annex A of Part 3 of Eurocode 8-Part 3 or Biskinis and
Fardis (2010) for other section types: (a) compact rectangular sections; (b) rectangular walls; (c) hollow or
anged rectangular members; (d) circular piers
4
Median:
u,exp = 0.99u,pred

12.5

3
u,exp: %

10
u,exp: %

Median: u,exp = 1.04u,pred

3.5

7.5

2.5
2
1.5

1
2.5
0.5
0

2.5

7.5
u,pred: %

10

12.5

2
u,pred: %

(a)

(b)

16
Median: u,exp = u,pred

14
Median:
u,exp = 1.035u,pred

12

5
u,exp: %

u,exp: %

10
4
3

8
6

1
0

2
0

3
4
u,pred: %

6
8
10
u,pred: %

(c)

12

14

16

(d)

where:
lm is the maximum of 0.4 m and the size of the bearing needed to support the vertical
reaction.


Leff
2 deg 2dg min 1;
(D6.35)
Lg
1

is the (static) relative displacement of the abutment and the (other support) of the deck
due to the spatially varying seismic ground displacement. dg in Eq. (D6.35) is the design
displacement of the ground from Eq. (D3.11) in Section 3.1.2.7 of this Guide. Leff is the
Figure 6.6. Minimum overlap length for seating of the deck on an abutment
lov

136

Chapter 6. Verication and detailing of bridge components for earthquake resistance

distance from the joint to the nearest point where the deck is considered not to move
horizontally with respect to the abutment in question: if a continuous deck is connected
with xity (including via seismic links or shock transmission units without force limiting
function) to more than one supports along its length, Leff is the distance to the centre of
this group of supports (including the special case of just one xed point all along this
length of deck); if it is supported everywhere on horizontally exible bearings, Leff is
measured to the centre of this group of bearings. Lg is the horizontal distance beyond
which the ground motion may be considered as fully uncorrelated (see Table 3.4 in this
Guide). If the bridge is less than 5 km from a known seismically active fault that can
produce an earthquake of magnitude M of at least 6.5, and unless a site-specic
seismological investigation is available, deg is taken as double that from Eq. (D6.35).
3 des is equal to the slack of any seismic links through which the deck may be connected to
piers or to an abutment, plus the total design displacement of the deck at the horizontal
level where it is seated on the abutment, due to the deformation of the structure in the
seismic design situation dEd:
dEd dE dG c2dT

(D6.36)

where dE is the design seismic displacement from Eq. (D5.48) in Section 5.9 (including any
effects of a torsional response about a vertical axis according to Section 5.7 of this Guide);
dG is the long-term displacement due to the permanent actions (including, for concrete
decks, prestressing after losses, shrinkage and creep); dT is the displacement due to the
design thermal actions; and c2 0.5 is the combination factor of thermal actions in the
seismic design situation in Table A2.1, A2.2 or A2.3 of EN 1990:2002.
The value of dEd at the horizontal level where the deck is seated on the abutment (i.e. at
the underdeck) is equal to the displacement at the level of the deck centroidal axis plus the
rotation of the deck end section multiplied by its centroidal distance from the underdeck.
For dE and dT this displacement and rotation should be taken with signs such that the
underdeck pulls away from the abutment (there is no choice for dG , which should be
calculated on the basis of the long-term value of any concrete creep, shrinkage and
prestressing). The design thermal actions considered for dT are the combination of (CEN,
2003a):
(a) The maximum contraction due to the difference, DTN,con , of the minimum uniform
bridge temperature component, Te,min , from the initial temperature, T0 , at the time the
deck is seated on the support. If T0 is not predictable, it is taken as the mean daytime
temperature during the foreseen construction period. Te,min is derived from the
characteristic value of the minimum annual shade air temperature, Tmin , at the site
according to the National Annex to EN 1991-1-5:2003 or as recommended by EN 19911-5:2003 itself in Figure 6.1 of EN 1991-1-5:2003. Tmin in turn is obtained from the
national isotherm maps at sea level in the National Annex to EN 1991-1-5:2003 and
adjusted for altitude and local conditions (e.g. frost pockets), for instance as
recommended in Annex A of EN 1991-1-5:2003.
(b) The vertical temperature difference component, DTM,heat , when the top surface of the
deck is hotter than the bottom one. Two alternative options are given in clauses 6.1.4.1
and 6.1.4.2 of EN 1991-1-5:2003 for its National Annex to choose one for the
determination of DTM,heat . Tables or gures with recommended values are also given
there.
The effects of the uniform and the temperature difference components are superimposed
linearly, but not in full. In symbolic terms, the combination recommended in EN 1991-15:2003 is the most adverse of
DTM,heat 0.35DTN,con ,

or

0.75DTM,heat DTN,con

(D6.37)

Albeit indirectly, Part 2 of Eurocode 8 makes reference for the above to the longitudinal direction
of the bridge, which is indeed the most critical for unseating. Regarding the transverse direction,
on the supply side the underdeck is usually quite wide; as far as the demand is concerned (i.e. the
right-hand-side of Eq. (D6.36)), dG and dT are normally zero, while the value of dE at the level of
the underdeck does not include a contribution from the rotation of the end section of the deck.
With these qualications, overlapping in the transverse direction may need to be checked only
137

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

under the narrow bottom ange of a prefabricated girder. In such a case, the torsional seismic
response about a vertical axis may have a non-negligible effect on the transverse displacement
of the end of the outer girder.

Clause 6.6.4(4) [2]

6.8.1.3 Overlap length at piers or at intermediate movement joints in the deck


Implicit in Section 6.8.1.2 is the assumption that the top of the abutment does not move. This
claim cannot be made for the top of an intermediate pier that supports the end of a deck
girder on bearings. If dEp denotes the design seismic displacement given by Eq. (D5.48) for the
top of the pier, Part 2 of Eurocode 8 makes the conservative assumption that dEp takes place
at the same time as the design seismic displacement of the end of the girder, dE, and out of
phase; so, it nds the minimum required overlapping of the end of the girder and the pier top
by adding dEp to the right-hand side of Eq. (D6.34). This applies, of course, both in the longitudinal and in the transverse directions.
At an intermediate movement joint in the deck within the span between adjacent piers or between
a pier and an abutment, the upper (supported) end of the deck on one side of the joint should
overlap the lower and supporting one by the square root of the sum of the squares of the
values obtained from Eq. (D6.34) for each one of these two deck ends. Note that such joints
are not very common in Europe. Part 2 of Eurocode 8 requires connection between the two
parts of the deck across the joint through seismic links (see Section 6.9).
6.8.1.4 Overlap lengths in skewed or curved bridges
Particular caution is needed if the axis of the deck is not at right angles to the line of support at
the abutments (skewed bridges), or if the two abutments are not parallel but at an angle to each
other.
If the bridge has a skew angle w . 0 (see Figure 5.12), the overlap length should be checked and
provided in the direction where the seat is narrowest; that is, at right angles to the edge of the deck
rather than in the longitudinal direction of the bridge. To this end, there are a few differences in
the way Eq. (D6.34) is elaborated:
lm  0.4 m still, but if there are bearings that are rectangular in plan, with sides bL and bT
in the longitudinal and the transverse directions, respectively, then also
lm  bL cos w bT sin w.
2 If the deck is connected to the abutment or the pier via seismic links having slack sL and
sT in the longitudinal and the transverse directions, respectively, the contribution of the
slack to des is sL cos w sT sin w.
3 If the displacements of the end of the deck due to permanent and quasi-permanent actions
are dG,L and dG,T in the longitudinal and the transverse directions, respectively, and the
corresponding values due to thermal actions are dT,L and dT,T (dG,T  0, dT,T  0), these
displacements enter into dEd in Eq. (D6.36) as dG dG,L cos w dG,T sin w  dG,L cos w,
and dT dT,L cos w dT,T sin w  dT,L cos w.
4 The design seismic displacement dE in Eq. (D6.36) is the peak value from Eq. (D5.48)
along a local axis at right angles to the edge of the deck; if no such local axis has been
introduced, the design seismic displacements dE,L and dE,T in the longitudinal and the
transverse directions, respectively, are projected onto the normal to the edge of the deck as
dE,L cos w and dE,T sin w, and combined to a single value through Eqs (D5.1) or (5.2).
1

Recall from Section 5.7.1 of this Guide the effects of twisting in skew bridges and the special
provisions in Part 2 of Eurocode 8 for their accidental eccentricities. If B and L are the width
and the length parallel to the sides in plan (see Figure 5.12), a twisting angle u produces at the
acute angle a displacement of u(B cos2 w L sin w)/2 normal to the edge of the deck. More
important, as pointed out in Section 5.7.1, if sin w . B/L, any twisting of the deck will
increase the joint width all along the width B of the support. To prevent long and narrow
decks with skew sin w . B/L from dropping by uncontrolled twisting, their seating at the
abutments may have to be much wider than what the calculations suggest.
If the axis of the bridge is at right angles to the abutments but is curved and the two abutments
are at an angle w to each other, the deck is again susceptible to dropping due to uncontrolled
138

Chapter 6. Verication and detailing of bridge components for earthquake resistance

twisting: if the projection of the outer corner of the deck (on the convex side) onto the line of
support at the other abutment falls outside the width B of the support (i.e. if the radius of
curvature at the axis is R . 0.5B(1 cos w)/(1  cos w), twisting of the deck will increase the
joint width all along B. Regarding the calculation of the overlap length with the help of
Eq. (D6.34), one should note that the abutments in the global longitudinal and transverse
directions are at an angle w/2 to the deck axis and the edge of the deck, respectively. So, it
would be convenient to apply Eqs (D6.34) and (D6.36) with displacement results from the
analysis in a local system of horizontal axes parallel and at right angles to the edge of the deck.
6.8.2
Clearances between the deck and critical or non-critical components
6.8.2.1 Clearance with critical components
Critical or major structural components should be protected from damage due to hard contact or
impact with the deck, by providing a clearance with it that can accommodate the total design
value of their relative displacement in the seismic design situation. The designer chooses which
components are sufciently critical or major to be fully protected from any damage through
this clearance. Apart from that, this clearance is also a means to avoid restraining the deck in
the seismic design situation in a way that introduces uncertainties in the response or complications of the analysis or design. For example, the clearance may be provided longitudinally
between the end of the deck and an abutment backwall that is not sacricial and not meant to
be engaged during the seismic response of the deck. It may also be provided transversely
between the side of the deck and a shear key that is meant to act as a second line of defence
against unseating in a seismic event beyond the design seismic action that may exhaust the
available overlap length or the displacement capacity of the bearings. The clearance does not
apply between the deck and a seismic link (e.g. a shear key) that is meant to be activated in
the seismic design situation. Any slack to be provided there is a design parameter.

Clauses 2.3.6.3(1)
2.3.6.3(4) [2]

The minimum value of the clearance is given by Eq. (D6.36), discounting the effects of the (static)
relative displacement of the component to be protected and the deck due to the spatially varying
seismic ground displacement (deg in Eqs (D6.34) and (D6.35)); dE , dG , c2 and dT are dened as in
Section 6.8.1.2, but only c2 has the same value, c2 0.5, because the way Eq. (D6.36) is applied
in Section 6.8.1.2 does not serve the purposes of the present section. In Section 6.8.1.2, all
displacement components refer to the horizontal level of the bearing; here, they are examined
at the two extreme horizontal levels where the deck and the component to be protected
overlap in the vertical direction the lowest and the highest (and computed as the deck
displacement at its centroidal axis plus the rotation of the deck end section multiplied by the
vertical distance of the centroid from the horizontal level considered). Second, apart from dG ,
which has a single sign and sense, the sense (sign) of all displacement components in Section
6.8.1.2 is taken such that the horizontal gap between the end of the deck and the supporting
component increases. Here, by contrast, the sense is the one that decreases the gap. Most
notable in this respect is the difference in the design thermal actions considered for dT . They
are the combination of:
(a) The maximum extension due to the difference, DTN,ext , of the maximum uniform bridge
temperature component, Te,max , from the initial temperature, T0 , at the time the deck is
erected, with Te,max derived from the characteristic value of the maximum annual shade air
temperature, Tmax , at the site according to the National Annex to EN 1991-1-5:2003 or as
recommended by EN 1991-1-5:2003 itself in Figure 6.1. Tmax is, in turn, read from the
national isotherm maps at sea level in the National Annex to EN 1991-1-5:2003, adjusted
for altitude and local conditions (e.g. as recommended in Annex A of EN 1991-1-5:2003).
(b) The vertical temperature difference component across the deck. For the displacement at
the highest horizontal level where the deck and the component to be protected overlap
vertically, this component is DTM,heat , when the top surface of the deck is hotter than the
bottom one; for the displacement at the lowest horizontal level of vertical overlapping, it is
DTM,cool when the deck top is cooler than its bottom. Tools or guidance for the
determination of DTM,cool and DTM,heat are given in clauses 6.1.4.1 and 6.1.4.2 of EN 19911-5:2003 and/or its National Annex.
The method recommended in EN 1991-1-5:2003 for the combination of the effects of the uniform
and the temperature difference components is similar to Eq. (D6.37); again in symbolic terms,
139

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

the most adverse of


DTM,heat 0.35DTN,ext

or

0.75DTM,heat DTN,ext

(D6.38a)

DTM,cool 0.35DTN,ext

or

0.75DTM,cool DTN,ext

(D6.38b)

If the component to be protected moves owing to the seismic action, its design seismic displacement relative to the deck, dE , may be taken as the square root of the sum of squares of the values
of the design seismic displacement calculated for each one of them at the horizontal level of
interest via Eq. (D5.48) in Section 5.9 (including any effects of torsional seismic response
about a vertical axis according to Section 5.7 of this Guide).
Needless to say, the clearance should be calculated and provided in the horizontal direction
where the deck and the component to be protected may come the closest.

Clause 2.3.6.3(5) [2]

6.8.2.2 Clearance with non-critical components


Repairable damage of non-critical components, such as replaceable roadway movement joints,
sacricial abutment backwalls, etc., due to the design seismic action is acceptable. For such
cases, Part 2 of Eurocode 8 reduces the required clearance between the deck and the noncritical component to a value that may be exhausted under a more frequent combination of
the seismic and the thermal actions (to be specied in the National Annex to Part 2 of
Eurocode 8). Its recommendation is for a clearance that will accommodate without damage
just 40% of the design seismic displacement, dE , plus all other displacement components in
Eq. (D6.36) in full; that is
dEd,oc 0.4dE dG c2dT

(D6.39)

where the sufx oc denotes occasional. Note that this level of protection is consistent with the
seismic action and performance requirements recommended for damage limitation in buildings
according to Part 1 of Eurocode 8 (CEN, 2004a).

6.9.
Clauses 6.6.1(2),
6.6.3.1(1), 6.6.3.1(2),
6.6.3.1(4), 6.6.3.1(5)
[2]

Seismic links

6.9.1
Denition and roles of seismic links
Part 2 of Eurocode 8 uses the general term seismic link for any special-purpose component
devised to transmit horizontal seismic forces from one part of the structure to another
usually from the deck to the piers or the abutments or across an intermediate movement joint
in a deck span without partaking in the vertical transfer of gravity loads. This differentiates
it from xed or elastomeric bearings, whose prime role is to transfer gravity loads from the
deck to the substructure, while sometimes doubling as means to transfer horizontal seismic
forces as well. Another feature of seismic links is that, although they are activated by the
horizontal seismic displacements of the bridge, they generally allow the non-seismic displacements to develop almost unrestrained.
Shear keys are usually employed as seismic links in the transverse direction, transferring horizontal forces across an intermediate separation joint in a deck span or from the deck to an abutment
or less often a pier, through compressive contact stress. Shear keys commonly have the form
of a short concrete corbel (sometimes prestressed), and are dimensioned and detailed as such.
Steel brackets are also used, if a proper connection detail is devised. Normally, non-seismic
displacements are zero in the transverse direction, and there is no need for a clearance
between the shear key and the part of the bridge from where it receives the horizontal force. A
vertical elastomeric bearing is often placed, though, between them, to avoid shock effects and
local damage upon hard contact, and to allow unrestrained relative rotation across the interface.
In the longitudinal direction, seismic links are normally steel linkage bars, bolts or cables,
connecting by tension the two parts of the deck across an intermediate movement joint in the
span, or the end of the deck to an abutment or pier, or the ends of two girders simply supported
over the same pier, etc. They normally have slack to allow unrestrained non-seismic displacements;
they are mobilised once the seismic displacements exhaust the slack. If the link is a rigid steel rod,
the slack may be provided between a steel plate at the end of the bar and the surface of the deck
component or abutment; a rubber ring or another device is provided along that gap for shock

140

Chapter 6. Verication and detailing of bridge components for earthquake resistance

absorption. Links connecting the parts of the deck across an intermediate movement joint in the
span serve only as second line of defence against unseating; they are not meant to be activated
in the seismic design situation. So, their slack should accommodate the relative displacement of
the two ends according to the following extension of Eq. (D6.38):
DdEd;12

q


2 d2 d
dE;1
G;1 dG;2 c2 dT;1 dT;2
E;2

D6:40

where the indices 1 and 2 refer to the ends of the two parts of the deck connected by the link. The
displacements due to the thermal action, dT,1 and dT,2 , are calculated for a uniform temperature
difference component causing contraction of the two linked parts of the deck and whatever
temperature difference component increases their distance at the horizontal level bridged by
the link. In symbolic terms, the combination of their effects is the most adverse of
DTM,heat 0.35DTN,con

or

0.75DTM,heat DTN,con

(D6.41a)

DTM,cool 0.35DTN,con

or

0.75DTM,cool DTN,con

(D6.41b)

As pointed out in Sections 4.5.4 and 5.5.1.1 of this Guide, activation of the link only after the gap
between the deck and the shear key closes or the slack of a tension link is exhausted is a nonlinear
feature of the behaviour, and should be taken into account in the modelling. As a minimum in the
context of linear analysis, Part 2 of Eurocode 8 requires the use of a linear spring with a stiffness
equal to the secant stiffness at yielding of the link; that is, equal to the yield force of the link
divided by the clearance or slack plus the elastic deformation of the link until it yields. It was
also pointed out that this is not very convenient for the analysis, because the link has not been
dimensioned yet at this stage of the design, and its yield force is not known. Of course, links
that are not meant to be activated in the seismic design situation do not need to be included in
the model.
6.9.2
Dimensioning of seismic links
As mentioned out in Section 6.3.2 (in point (b)), a seismic link should be dimensioned for
capacity design effects. In bridges of limited ductile behaviour, these may be taken as the
seismic action effects in the link from the linear analysis multiplied by q. In bridges of ductile behaviour, the capacity design effects should be determined through Eq. (D6.8) and the associated
procedure highlighted in Section 6.4.2; the simplication of Eq. (D6.11) may be employed to this
end.

Clauses 2.3.6.2(2),
5.3(2), 6.6.3.1(3) [2]

As an exception to the general rule above, in seismic links connecting in tension the two parts of
the deck across an intermediate movement joint in the span, or the end of the deck to an
abutment or pier, or the ends of two girders simply supported over the same pier, etc., the
linkage element may be dimensioned for (the results of ) a longitudinal force equal to
1.5agSMd , where ag is the design ground acceleration on type A ground, S is the soil factor
and Md is the mass of the (part of the) deck linked to the pier or abutment, or the least of the
masses of the two parts of the deck on either side of the intermediate separation joint. Alternatively, the longitudinal force for the design may be obtained from a more accurate analysis,
taking into account the dynamic interaction of the parts of the deck connected by the link.
Once the capacity design internal forces are established, the seismic link is dimensioned at the
ULS according to the relevant material Eurocode: a concrete shear key as a corbel according
to EN1992-1-1:2004, a steel bracket or a linkage rod, bolt or cable according to the relevant
part of Eurocode 3. The design shear resistance of concrete shear keys is divided by a Nationally
Determined Parameter reduction factor gBd . Part 2 of Eurocode 8 recommends a value of
gBd 1.25 for bridges of limited ductile behaviour. For those of ductile behaviour, it is
allowed either:
g
g

Clauses 5.6.2(2)b,
5.6.3.3(1)b [2]

to take the same value as for bridges of limited ductile behaviour or


to subtract from it qVEd/VC,o  1 (but not to a nal value below gBd 1), where:
VEd is the maximum shear force of the seismic link from the analysis for the seismic
design situation
VC,o the capacity design shear of the link, without the upper limit VC,o  qVEd .
141

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

(see also notes (4) and (5) in Table 6.1). Note that, as the behaviour of any seismic link is normally
non-ductile, it is prudent to reduce its design resistance by the factor gBd specied in Part 2 of
Eurocode 8 for concrete shear keys, although Part 2 of Eurocode 8 does not require it explicitly.

6.10.

Dimensioning of bearings

6.10.1 Introduction
Sections 6.10.2 to 6.10.4 refer to bearings supporting the deck over certain piers and/or at the
abutments, while other supports (usually piers) are connected to the deck monolithically or
via xed bearings. Movable bearings and other special devices (isolators) arranged along a
continuous interface between the underdeck and all its supports (isolation interface) are dealt
with in Chapter 7.
Clause 7.5.2.3.5(5) [2] Part 2 of Eurocode 8 also allows the use of common at sliding bearings (without a controlled

lower limit of friction coefcient) and/or common low-damping elastomeric bearings at the
isolation interface of bridges with seismic isolation. Bearings for this use are a special case,
covered in this section. Conformity with EN 1337-2 (CEN, 2000) and EN 1337-3 (CEN,
2005d), respectively, and with certain additional design requirements specied in Part 2 of
Eurocode 8 is sufcient for them.
6.10.2

Fixed bearings

Clauses 6.6.2.1(1),
As mentioned in Section 6.3.2 (in point (b)), xed bearings should be dimensioned in the seismic
6.6.2.1(2), 6.6.3.2(2)b, design situation for capacity design effects. In bridges of limited ductile behaviour this just entails
6.6.3.2(3) [2]
multiplying by q the seismic action effects from a linear analysis. By contrast, in bridges of ductile

behaviour, the general procedure of Section 6.4.2 and Eq. (D6.8) should be used, possibly with
the simplication of Eq. (D6.11). As an exception, Part 2 of Eurocode 8 allows xed bearings to
be dimensioned using only the analysis results for the seismic design situation, provided they can
be easily replaced and are supplemented with seismic links that are capacity-designed to provide
instead the required horizontal resistance (i.e. dimensioned for the horizontal resistance of the
xed bearings with capacity design amplication).
Fixed bearings are normally non-ductile. Although Part 2 of Eurocode 8 does not require it explicitly, it is prudent to reduce their design horizontal resistance by the factor gBd mentioned at the
end of Section 6.9.2 for the design shear resistance of concrete shear keys.

Clauses 5.3(7),
7.6.2(1) [2]

6.10.3 Flat sliding bearings


Section 6.4.5 pointed out that the elements supporting the deck of ductile bridges should be
designed for a horizontal friction force in sliding bearings equal to gofRdf , where gof 1.30
and Rdf is the maximum design force in the seismic design situation. The dimensions of the
sliding plate of the bearing should be sufcient to accommodate, with adequate safety margin,
the extreme design displacement in the design seismic situation, dEd . The value of dEd is given
by Eq. (D6.36), and computed at the location and the horizontal level of the bearing according
to Section 6.8.1.2, except that, if the sliding bearing belongs to a seismic isolation system, the
value dE,a gISdE , is used in lieu of the seismic displacement from the analysis, dE , with a
value of 1.50 recommended in Part 2 of Eurocode 8 for gIS . This value may be considered to
provide the reasonable safety margin for such bearings. For those not belonging to such a
system, and in the absence of more specic guidance from Part 2 of Eurocode 8, this margin
may be taken to be between 20% and 30% of dE , depending also on the sensitivity of the
estimation of dE on the value of (EI )eff used in the analysis.

6.10.4 Simple low-damping elastomeric bearings


6.10.4.1 Scope and denitions
Clauses 6.6.2.3(1), 7.2, An elastomeric bearing is a block of vulcanised elastomer with natural or chloroprene rubber as
7.5.2.3.3(1),
the raw polymer reinforced internally with steel plates, chemically bonded to the elastomer during
7.5.2.3.3(2),
vulcanisation. Being nearly incompressible, the rubber exhibits signicant lateral expansion when
7.5.2.3.3(4) [2]
in vertical compression. The steel plates restrain this expansion through shear stresses at the
interface with the rubber. These stresses are largest at the perimeter of the steel plate, and may
cause debonding there. The magnitude of these stresses is usually checked through the associated
shear strains, g. The large magnitude of these shear strains is manifested by the bulging of each
elastomer layer, which is largest at the perimeter and right next to the steel plate.
142

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Part 2 of Eurocode 8 distinguishes two types of elastomeric bearing, depending on the damping
they offer under cyclic shear. The low-damping ones exhibit narrow hysteresis loops with an
equivalent viscous damping ratio, j, of less than 6%. Accordingly, their shear behaviour may
be approximated as linear elastic and characterised by the shear modulus of the elastomer
alone, G, which applies throughout the total thickness of the elastomer in the bearing (tq in
Eq. (D6.43b)). The damping they offer is consistent with the default value of 5% used in
linear analysis, as well as the baseline for the q factors adopted in Part 2 of Eurocode 8.
If the deck is not xed to any pier or abutment (not even through seismic links) and is seated on
elastomeric bearings at all of its supports (or possibly on at sliding bearings over certain
supports), therefore forming an isolation interface between the deck and all of its supports,
the full horizontal seismic force of the deck is transmitted through these elastomeric bearings.
Such bearings have a decisive role in the global response of the bridge to the horizontal
seismic components. Part 2 of Eurocode 8 acknowledges them as seismic isolation devices, and
considers the deck as seismically isolated. It calls them simple low-damping elastomeric
bearings, dened as laminated low-damping elastomeric bearings in accordance with EN 13373:2005, not subject to EN 15129:2009 (Antiseismic Devices) (CEN, 2009) nor to any special
tests for seismic performance. Accordingly, they are mainly covered in Section 7 of Part 2 of
Eurocode 8, dedicated to bridges with seismic isolation. The increased reliability required
from the isolation system is implemented in that case by increasing the design seismic displacements to dE,a gISdE , with a value of 1.50 recommended in Part 2 of Eurocode 8 for gIS . An
additional requirement is a detailed explicit investigation of the inuence of the variability of
the design properties of the isolators on the seismic response.

Clauses 6.6.2.3(1)
6.6.2.3(4),
7.5.2.3.3(5),
7.5.2.3.3(6),
7.5.2.4(1), 7.5.2.4(5),
7.5.2.4(6), 7.6.2(1),
7.6.2(2), 7.6.2(5) [2]

If the deck is xed to the top of one or more piers (or an abutment), directly or via seismic links,
the seismic action is transferred to the substructure through this type of connection. Any elastomeric bearings used over other supports of the deck have a local role in the overall seismic
response of the bridge. Such bearings are designed, according to EN 1337-3:2005 (CEN,
2005d), to resist all non-seismic horizontal and vertical actions. They are also designed according
to Part 2 of Eurocode 8 to accommodate the imposed deformations due to the design displacement in the seismic design situation, as given by Eq. (D6.36) with the value dE as the design
seismic displacement, without multiplication by gIS . A detailed investigation of the variability
of the design properties of such bearings is not required, in contrast to bearings used as
isolator units. Summing up, with these differentiations, low-damping elastomeric bearings subjected to seismic shear deformations are veried in the seismic design situation with the rules
given in EN 1337-3:2005 and complemented in Part 2 of Eurocode 8.
The following section covers the design and verication of simple low-damping elastomeric
bearings according to EN 1337-3:2005 and Part 2 of Eurocode 8, differentiating wherever
pertinent between seismic and non-seismic design situations. Design with/of high-damping
elastomeric bearings is addressed in Chapter 7 of this Guide. In the remainder of this section,
simple low-damping elastomeric bearings in the sense of Part 2 of Eurocode 8 are termed, for
brevity, elastomeric bearings or even just bearings.
Elastomeric bearings are usually circular, with diameter D, or rectangular, their sides denoted
here as bx and by; the longer side is normally in the transverse direction of the bridge, to
minimise the rotational restraint in the longitudinal direction and if used under a prefabricated girder to stabilise it laterally during erection. The effective plan dimensions of the
bearing are considered to be those of its steel reinforcing plates, denoted here as D0 , bx0 , by0 .
According to EN 1337-3:2005, at least 4 mm of elastomer should cover laterally the edge of
the steel plate:
D0  D  8

bx0  bx  8

by0  by  8

(dimensions in mm)

(D6.42)

The effective plan area of the bearing, A0 , is determined from its effective plan dimensions
above.
If the bearing has n internal layers of elastomer with thickness ti (25 mm  ti  5 mm according to
EN 1337-3:2005) and n 1 steel plates of thickness ts (ts  2 mm according to EN 1337-3:2005),
143

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

and the top and bottom steel plates have an outer cover of elastomer, at least 2.5 mm thick per
EN 1337-3:2005, the total nominal thickness of the bearing is
tb nti (n 1)ts 5

(dimensions in mm)

(D6.43a)

and that of the elastomeric part is


tq nti 5

(dimensions in mm)

(D6.43b)

Only the elastomer is deformable in shear. The nominal value of the shear modulus of elastomeric
bearings, as well as its value determined by testing, denoted in EN 1337-3:2005 as G and Gg ,
respectively, refer to the total elastomer thickness, tq . EN 1337-3:2005 species a target mean
value of 0.9 MPa for Gg (and allows the designer to specify alternative ones of 0.7 MPa or
1.15 MPa), with an acceptable test value tolerance of +0.15 MPa (if one of the two alternatives
is specied, +0.1 MPa or +0.2 MPa, respectively). Accelerated ageing (3 days at 708C) may not
increase the value of Gg by more than 0.15 MPa. Note that, according to Annex F of EN 13373:2005, the value of Gg is measured under normal temperatures between shear strains gq,d 0.27
and 0.58. Such values are representative of the normal-temperature stiffness of bearings in the
strain range allowed under non-seismic design situations (gq,d  1.0). In the signicantly wider
strain range allowed in the seismic design situation (gq,sd  2.0), substantially higher values of
the secant shear modulus at peak strain are expected. For this reason, Part 2 of Eurocode 8
gives a higher value Gb aGg for the nominal design value of the shear modulus of elastomeric
bearings with the value of a (normally in the range 1.11.4) determined from tests. This value is
used as lower-bound design property (LBDP), if the bearing is used as isolator unit, with a
value of 1.2Gb as the upper-bound design property (UBDP) for the minimum bearing temperature
for seismic design, Tmin,b , not less than 08C, or obtained from the appropriate l factor from
Annex JJ of Part 2 of Eurocode 8 if Tmin,b is less than 08C.
Among the various types of elastomeric bearings specied in EN 1337-3:2005, two deserve special
mention: type C, with top and bottom steel plates (proled or allowing xing to the parts of the
bridge on either side of the bearing), and type B, without steel or any other type of plate attached
to the top or bottom surface of the elastomer. If 8 mm  ti , steel plates of type C bearings
should be at least 15 mm thick; if ti . 8 mm, their minimum thickness is 18 mm (CEN,
2005d). EN 1337-3:2005 lists recommended standard sizes for type B bearings (which are also
the basis for type C ones). As an indication of the range:
g
g
g
g

bearings from 100  150 mm (or 1200 mm) to 250  400 mm (or 1350 mm) have n 27
elastomer layers with ti 8 mm and steel plates with ts 3 mm
for bearings of 300  400 mm (or 1400 mm) to 500  600 mm (or 1650 mm), n 310,
ti 12 mm, ts 4 mm
for bearings of 600  600 mm (or 1700 mm) to 700  800 mm (or 1850 mm), n 410,
ti 16 mm, ts 5 mm
for bearings of 800  800 mm to 900  900 mm (or 1900 mm), n 411, ti 20 mm,
ts 5 mm.

In these recommended sizes, the ratio of the mean horizontal dimension to the total thickness, tb ,
ranges from about 3 to about 9.
The shape factor, S, is the effective plan area of a bearing, A0 , divided by the product of the
perimeter dened by its effective plan dimensions and the thickness of an inner elastomer
layer:
g

for rectangular bearings


S bx0 by0 /[2(bx0 by0 )ti]

for circular bearings


S D0 /4ti

144

(D6.44a)

(D6.44b)

Chapter 6. Verication and detailing of bridge components for earthquake resistance

If the top of the bearing is displaced with respect to the bottom by ddx  0 in the horizontal
direction x and by ddy  0 along the orthogonal direction, y (for rectangular bearings, along
the sides), the vertical load passes through the area in plan where the displaced top and
bottom overlap. This is termed the reduced effective plan area, and denoted as Ar:
g

for rectangular bearings


Ar (bx0  ddx)(by0  ddy)  (1  ddx/bx0  ddy/by0 )A0

(D6.45a)

for circular ones (Constantinou et al., 2011)


Ar A0 (d  sin d)/p

(D6.45b)

where

d 2 cos  1(dd/D0 ) (in rad)

(D6.46)

with
dd

q
2 d2
ddx
dy

D6:47

6.10.4.2 Dimensioning of elastomeric bearings according to EN 1337-3:2005 and


Eurocode 8
6.10.4.2.1 Verification of shear strains
EN 1337-3:2005 denes the shear strain in the elastomer due to the total design displacement dd
dened via Eq. (D6.47) as

gq,d dd/tq

(D6.48)

where tq is the total thickness of the elastomer from Eq. (D6.43b). It further limits the value of
gq,d at the ULS of the bearing under the total design displacement induced by the fundamental
combination of actions in EN 1990:2002, to a maximum value of 1.0 (see Section 6.10.4.3 for
details about the fundamental combination and the horizontal displacements it induces).

Clauses 6.6.2.3(1),
6.6.2.3(2), 6.6.2.3(4),
7.5.2.3.3(5),
7.5.2.3.3(6), 7.6.2(1),
7.6.2(2), 7.6.2(5)
7.6.2(7) [2]

For low-damping elastomeric bearings veried in the seismic design situation, whether they are
used as isolators or not, Part 2 of Eurocode 8 imposes the following additional requirement on
the maximum total design strain in the seismic design situation, gb,Ed:

gb,Ed  2.0

(D6.49)

where gb,Ed is determined according to Eqs (D6.47) and (D6.48) from the design horizontal
displacements in the seismic design situation, as given by Eq. (D6.36). If the bearing is part of
the isolation system, the value dE,a gISdE is used as design seismic displacement, with a value
of 1.50 recommended in Part 2 of Eurocode 8 for gIS . This denition of the maximum total
design strain in the seismic design situation, gb,Ed , applies also in all other verications specied
in EN 1337-3:2005 for elastomeric bearings, wherever bearing shear strains in the seismic design
situation are involved.
Equation (D6.49) is the only additional verication specied by Part 2 of Eurocode 8 for the
seismic design situation in addition to those of EN 1337-3:2005. Equations (D6.43), (D6.47) and
(D6.48) or (D6.49) readily yield a rst estimate of the number of internal elastomer layers, n.
If the elastomeric bearings constitute the restoring element of an isolation system, Eq. (D6.49)
usually controls the elastomer thickness in the bearing, and often its plan dimensions.
Note that Section 7 of Part 2 of Eurocode 8 requires all verications of elements of the isolation
system, the superstructure and the substructure be carried out for the most adverse results of two
analyses: one based on the UBDP and another on the LBDP of the isolating units.
145

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

6.10.4.2.2 Verification of buckling


The buckling load of a bearing without lateral displacement dd is (Constantinou et al., 2011)
Ncr lA0 r0

pGS
nti

D6:50

p
where in a rectangular bearing l  1.5 and in a circular one l  2 (Constantinou et al., 2011), A0
p
is the effective plan area of the bearing, and r0 (I0 /A0 ) its minimum radius of gyration, G is the
nominal shear modulus of the elastomer and S is the shape factor from Eqs (D6.44). Therefore:
g

for a rectangular bearing with bx0  by0


Ncr 0:68A0

0
b02
x by G
nt2i b0x b0y

D6:51a

for a circular one


Ncr A0

D02 G
3:6nt2i

D6:51b

If the bearing is bolted at its top and bottom to mounting plates, Eqs (D6.51) apply in good
approximation in the laterally deformed conguration, using there the reduced effective plan
area Ar from Eqs (D6.45) in lieu of A0 . If it is dowelled to them or kept in recessed plates, the
lateral displacement may roll the bearing over, in which case Eqs (D6.51) are not critical. As
in the fundamental combination of actions in EN 1990:2002, to which EN 1337-3:2005 refers
for the buckling stability of the bearing at the ULS, the lateral displacements are rarely large,
EN 1337-3:2005 does not consider the laterally deformed conguration of not bolted bearings
as a separate case, and checks the buckling stability of the bearing on the basis of Eqs (D6.51)
with a safety factor of about 2 for rectangular bearings or of 1/0.6 for circular bearings:
g

for a rectangular bearing with bx0  by0


0
b02
Nd 2b0x GS
x by G


Ar
3nti
3nt2i b0x b0y

D6:52a

for a circular bearing


Nd 2D0 GS D02 G


Ar
3nti
6nt2i

D6:52b

If both ddx and ddy are signicant in magnitude, Eq. (D6.52a) reduces to a quadric equation in bx0
for given bx0 /by0 . In approximation:
1
v
0
s
u


2
u
0
0
1 ddx ddy t ddx ddy
b 3nNd
b
D6:53a

1 x0 A
b0x  @
4ti x0
2
2
by
G
by
2
If dd is non-negligible, Eq. (D6.52b) is strongly nonlinear in D0 ; the required bearing diameter
may be found iteratively from Eq. (D6.46) and
s
4
24nt2i Nd
D0 
Gd  sin d

D6:53b

The above verication of the buckling load of the bearing should be carried out for the fundamental combination of actions in EN 1990:2002, as well as in the seismic design situation with
the relative displacements ddx , ddy (or dd from Eq. (D6.47)) corresponding to the total design
displacement according to Eq. (D6.36), with the value dE,a gISdE used as the design seismic
displacement and a value of 1.50 recommended in Part 2 of Eurocode 8 for gIS , if the bearing
146

Chapter 6. Verication and detailing of bridge components for earthquake resistance

is part of the isolation system. However, as this verication is controlled by the design vertical
force on the bearing, Nd , whose value in the seismic design situation is signicantly smaller
than in the fundamental combination of actions in EN 1990:2002, the seismic design situation
is rarely critical.
6.10.4.2.3 Verification of the total shear strain, gt,d
As pointed out in the opening paragraph of Section 6.10.4.1, the elastomer may debond from the
steel plate, owing to large shear strains, g, at their interface. The maximum value g in the
elastomer is

gt,d gq,d gc,d ga,d

(D6.54)

where:
g
g

gq,d is the shear strain in the elastomer due to the total design displacement, dened in
Eq. (D6.48), or in Eq. (D6.49) as gb,Ed for the seismic design situation.
gc,d is the maximum shear strain in the elastomer due to the design value of the vertical
compression, Nd; it takes place at the interface with the steel plate at the mid-side of
rectangular bearings or all around the perimeter in circular ones, and is equal to
gc;d

f1 N d
Ar GS

D6:55

with G, Ar and S as for Eq. (D6.50). Values of f1 are tabulated in Constantinou et al.
(2011) as a function of the geometry of the bearing and of the ratio of the bulk modulus of
the elastomer, K (K  2000 MPa) to its shear modulus G, with an upper bound for circular
bearings of
f1  1 2

S2 G
K

D6:56

Approximate expressions are also proposed for f1 in Stanton et al. (2008), including a
value f1 1.0 for circular bearings. EN 1337-3:2005 xes f1 to
f1 1.5

(D6.57)

which is on the safe side for the G values specied and the bearing sizes recommended in
EN 1337-3:2005, especially for circular and square bearings.
ga,d is the maximum shear strain in the elastomer due to the total design angular rotations
in the vertical planes through axes x and y, adx  0, ady  0, respectively; it takes place at
the extreme compression bres of the bearing in the vertical plane of the rotation and is
equal to (Constantinou et al., 2011)
for rectangular bearings

ga;d f2

adx b2x ady b2y


nt2i

D6:58a

for circular bearings

ga;d f2

ad D02
nt2i

D6:58b

where

ad

q
a2dx a2dy

D6:59

Constantinou et al. (2011) tabulate values of f2 as a function of the geometry of the


bearing and the K/G ratio of the elastomer, while Stanton et al. (2008) propose
147

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

approximate expressions for it, including a value f2 3/8 for circular bearings. EN 13373:2005 xes f2 to
f2 0.5

(D6.60)

which is again on the safe side for the G values specied and the bearing sizes
recommended in EN 1337-3:2005, especially so for circular and square bearings.
According to EN 1337-3:2005, the value of gt,d from Eqs (D6.48), (D6.54), (D6.55) and (D6.57)
(D6.60) for the fundamental combination of actions in EN 1990:2002 should be limited upwards
to 7/gm , with gm 1.0. Part 2 of Eurocode 8 also recommends the same upper limit (i.e. with
gm 1.0) to the value of gt,d computed from Eqs (D6.49), (D6.54), (D6.55) and (D6.57)
(D6.60) for the seismic design situation. The upper limit on gt,d is translated to additional
constraints on the plan dimensions or the number of elastomer layers, n. The latter is easier to
implement:
g

for rectangular bearings


02
02
dd adx bx ady by

ti
2t2
 i0

n
3ti Nd bx b0y
7
 
 

gm G b0x b0y 2 1  ddx =b0x  ddy =b0y

D6:61a

for circular ones


dd ad D02

ti
2t2i
n
7
24ti Nd

03
gm GD d  sin d

D6:61b

with dd from Eq. (D6.47). For circular bearings, ad is given by Eq. (D6.59) and d from
Eq. (D6.46).
Note that the number of layers, n, comes out of Eqs (D6.48) and (D6.49) as independent of the
size of the bearing in plan. According to Eqs (D6.52) and (D6.53), the taller the bearing, the
larger its plan dimensions should be against buckling under vertical compression. By contrast,
Eqs (D6.61) give fewer layers if the size in plan increases. If the second estimate of n from Eqs
(D6.61) is larger than that from Eqs (D6.48) and (D6.49), Eqs (D6.53) may have to be revisited;
if in that case they suggest a larger size in plan, it is the turn of Eqs (D6.60) to be revisited, until a
stable combination of elastomer thickness and plan dimensions is reached. Note that, if the
number of bearings sharing the vertical reaction increases, the required plan area per bearing
drops due to the reduction in Nd , but the total area increases (see Eqs (D6.53): as the horizontal
displacements of the bearing are independent of its size in plan, the reduction in the effective plan
area according to Eqs (D6.45) is proportionally larger in smaller bearings). Unless it is controlled
by Eqs (D6.48) or (D6.49), the number of elastomer layers may decrease as well.
6.10.4.2.4 Fixing of elastomeric bearings
If the bearing is not positively xed to both the underdeck and the top of the pier or the
abutment, (e.g. for type B bearings), EN 1337-3:2005 requires checking of the transfer of the
design shear force VEd at the ULS through friction:

148

Nd;min
MPa  3
Ar

D6:62a

Vd
b
 0:1
Nd;min
Nd;min
MPa
Ar

D6:62b

Chapter 6. Verication and detailing of bridge components for earthquake resistance

where
g
g

b 0.9 for bearings on concrete


b 0.3 for bearings on all other types of surface (metallic, bedding resin mortar, etc.).

Vd and Nd,min are the design shear force and the corresponding minimum axial force, respectively, on the bearing, and Ar is given by Eqs (D6.45). Equations (D6.62) are to be met:
g

under the fundamental combination of actions in EN 1990:2002, with the partial factors
applied on span permanent actions so that they produce the minimum vertical
compression on the bearing and with the factored trafc loads arranged along the deck so
that they give the maximum vertical tension
in the seismic design situation, with the seismic action causing the maximum possible
bearing tension and the maximum horizontal displacements, dsd,x , dsd,y , that are consistent
with the design shear force, Vd .

The ratio Vd/Nd,min in Eq. (D6.62b) is dened in EN 1337-3:2005 as the friction coefcient mf .
Although not explicit, such friction refers to the interface between the external elastomeric
layer of a type B bearing and the material of the underdeck, the pier or the abutment in
contact with it. Therefore, Eq. (D6.62b) does not apply if the external steel plates of other
types of elastomeric bearings are appropriately bonded to the bridge elements supported on
the bearing or supporting it.
If Eqs (D6.62) are not met, the bearing should be xed at its top and bottom. The end plates
(those of type C bearings of EN 1337-3:2005) should be bolted or dowelled to the top and
bottom mounting plates or kept in place in recesses of the mounting plates. Note that this
way of xing the bearing also prevents it from slipping from its intended position during
erection and if it is xed by bolting precludes partial uplift due to the combination of
lateral displacements and rotation, which will increase the shear strains on the compressed
side over the predictions of Eqs (D6.54)(D6.60), and may even cause the bearing to roll
over.
6.10.4.3 Action effects for the verication of simple low-damping elastomeric
bearings
Elastomeric bearings should meet the requirements in EN 1337-3:2005 for non-seismic actions,
notably for the fundamental combination in EN 1990:2002, Eq. (6.10),
X
X
gG; j Gk; j gP P gQ;1 Qk;1
gQ;i c0;i Qk;i
j1

i>1

or Eq. (6.10a),
X
X
gG; j Gk; j gP P gQ;1 c0;1 Qk;1
gQ;i c0;i Qk;i
j1

i>1

or Eq. (6.10b),
X
j1

jj gG; j Gk; j gP P gQ;1 Qk;1

gQ;i c0;i Qk;i

i>1

where in this case Gk are the permanent actions, Pk is prestressing after losses, Qk,1 is trafc loads
with their characteristic value and Qk,i (i . 1) is the characteristic value of thermal actions, Tk,i .
The choice between Eq. (6.10), (6.10a) or (6.10b) in EN 1990:2002 as well as the partial or combination factors gG, j , gQ,1 , gP , j or c0,1 are National Determined Parameters. The reader is
referred to EN 1990:2002 for the recommended choices, those of gQ,1 and c0,1 depending on
the type of trafc load and model. The recommended value of the product gQ,ic0,i for the
thermal actions is gQ,ic0,i 0.9. In what follows, the fundamental combination is symbolically
written as gGP gQ 0.9Tk , with gGP representing the part of the fundamental combination
due to the permanent actions Gk and Pk with the appropriate values of gG,j , gP and j applied, gQ
149

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

that due to the trafc actions Qk,1 with the associated values of gQ,1 , c0,1 and Tk the characteristic
thermal actions. The total effects of the fundamental combination are indexed by FC, and the
individual ones due to gGP, gQ and Tk as dened above by gGP, gQ and T, respectively.
Note that gQ,1 0 over the plan area of the deck whose loading by trafc is favourable. By
the same token, gG, j 1.0 and j 1.0 over the length of the deck whose loading by permanent
actions is relieving; however, if the action effects are not very sensitive to variations in the
permanent actions along the length of the deck (and in the present case they are not), constant
values of gG, j and j may be used throughout the deck.
Equation (D6.48) should be checked for the maximum horizontal displacements ddx , ddy of the
top of the bearing relative to the bottom due to the fundamental combination of actions. By
contrast, Eqs (D6.52), (D6.53) and (D6.62b) involve both Nd and ddx , ddy, , making it necessary
to search for an arrangement of gravity loads along the bridge that is most adverse for the verication of interest. The relative horizontal displacement of the bearing due to the fundamental
combination of actions is
dFC dgGP dgQ 0.9dT

(D6.63)

where dgGP is the long-term relative displacement due to the factored permanent actions
abbreviated above as gGP, including prestressing after losses, shrinkage and creep, if relevant,
(no partial factor with value greater than 1.0 is applied on deformations due to concrete
shrinkage); dgQ is the short-term relative displacement due to the factored trafc actions
denoted above as gQ; and dT is the relative displacement due to the characteristic thermal
actions.
These displacement components are normally very small (almost zero) in the transverse direction.
Hence, in the following they are assumed to be in the longitudinal direction. They all refer to the
horizontal level where the deck is seated on the bearing (i.e. the underdeck), and are equal to the
displacement at the level of the centroidal axis of the deck plus the rotation of the deck section
above the bearing multiplied by the centroidal distance to the underdeck. As it is the absolute
value of horizontal displacements that enters into the verications, in order for dFC to have
the largest possible absolute value, dgQ and dT should have the (xed) sign of dgGP:
g

Owing to concrete creep, shrinkage and prestressing, in concrete decks dgGP is towards the
mid-length of the deck and away from its nearest end. Then, dgQ and dT should also be
towards the deck mid-length. For dgQ , this means that the arrangement of trafc loads
along the bridge should be that giving the maximum hogging moment (negative) at the
span next to the support in question on the side towards the deck mid-length. (If the
values of gG, j and j are differentiated along the deck, they are taken as gG, j 1.0 and
j 1.0 in the span next to the support of interest on the side towards the deck mid-length
and with their non-unity values in the spans to its left and right, alternating with the unity
values thereafter.) The thermal displacement, dT , is due to the same thermal actions
highlighted in Section 6.8.1.2 in points (a) and (b). Indeed, if a value of dT at a bearing is
computed according to Section 6.8.1.2 for the purposes of overlap length, it applies for the
purposes of Eq. (D6.57) as well.
Under a non-concrete deck, dgGP may be in the same direction as above, or in the reverse
direction (i.e. away from the deck mid-length and towards its nearest end). In this latter
case, the reverse also applies regarding the most unfavourable longitudinal arrangement of
trafc loads and of the spans where gG, j and j assume unity or non-unity values; as far as
dT is concerned, it is the same as the value computed in Section 6.8.2.1 in points (a) and
(b) for the purposes of clearance at the lowest horizontal level (in which case Eq. (D6.38b)
applies).

The values of ddx and ddy are quite sensitive to the thermal actions but insensitive to those due to
trafc that determine dgQ . The reverse holds for Nd . So, when the maximum value of Nd is of
interest, the position of trafc loads (and the spans where gG, j and j may assume their nonunity values) are those producing the maximum vertical reaction on the bearing. By contrast,
the thermal actions are chosen to maximise the concurrent absolute values of ddx and ddy (as
they always affect unfavourably the verication, see Eqs (D6.53), (D6.61) and (D6.62)).
150

Chapter 6. Verication and detailing of bridge components for earthquake resistance

When Eqs (D6.61) are veried in the seismic design situation, the value of Nd is derived from the
condition when the seismic actions induce compression in the bearing. The displacements ddx and
ddy are those associated with Eq. (D6.49) in Section 6.10.4.2.1 and specied there. If the deck is of
concrete, the longitudinal displacements are computed as in Section 6.8.1.2, with movement
towards the deck mid-length and away from its nearest end. For other deck materials, the
reverse direction of longitudinal movement may need to be examined in addition, with displacements computed as in Section 6.8.2.1 in points (a) and (b) for the clearance at the lowest horizontal level (with Eq. (D6.38b) applying). It is meaningful to check Eqs (D6.49), (D6.53), (D6.61) or
(D6.62) in the transverse direction only when the design seismic displacement of the bearing in
that direction is markedly larger than in the longitudinal direction, offsetting the almost zero
values of dG and dT .
6.10.4.4 Sizing of simple low-damping elastomeric bearings for seismic isolation
6.10.4.4.1 Introduction
If the elastomeric bearings resist alone the seismic action (considered in Part 2 of Eurocode 8 as
seismic isolators), the elastomer thickness and the bearing dimensions in plan determine the
lateral stiffness and the seismic displacement and force demands. So, the thickness cannot be
obtained from the displacement through simple application of Eq. (D6.49). In the following, a
procedure is proposed for the preliminary estimation of the thickness and the dimensions in
plan of the bearing. It assumes that the exibility of the piers and the abutments may be
ignored compared with that of the bearings, and that all elastomeric bearings have the same
total elastomer thickness, tq . For such a system the rigid deck model may be applied. It is
further assumed that the centre of the horizontal bearing stiffnesses coincides with the centre
of mass of the
P deck, so that no twisting response occurs. The total plan area of the bearings is
denoted by
Ab . The total design vertical load of the deck,Pcorresponding to the total deck
tributary mass in the seismic design situation, is denoted as
Nd . Further, we introduce the
notation:
Sag
(design ground acceleration at the top of the soil in g
g
P
N
sb P d
Ab
a

sb
Gb

D6:64a
D6:64b
D6:64c

P
Witness that, since
Nd is considered as given,P
sb and its dimensionless counterpart r are inverbearings that undertake
sely proportional to the sum of the plan areas,
Ab , of the elastomeric
P
the entire horizontal seismic force. Note also that, even though
Nd is carried by the same elastomeric bearings as this horizontal force,
P sb is not necessarily equal to the normal stress of each
bearing, because the distribution of
Nd to the individual bearings depends on various factors,
unlike that of the horizontal force, which is distributed in proportion to thePplan area of each
bearing. Note also that sb and r can be substantially varied by varying 1/ Ab , not only by
changing the areas of individual bearings: sb and r may increase by assigning part of the
gravity load to a number of at sliding bearings, reducing thereby the elastic restoring force of
the system.
6.10.4.4.2 Seismic demand for the bearing shear strain
Total horizontal stiffness:
Kb

Gb

Ab

tq

D6:65

Period:
s
s

P
r
r
P
rt q
Nd tq
Nd
sb t q
P
T 2p
2p
2p
2p
gKb
gGb Ab
Gb g
g

D6:66

151

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Seismic shear strain of bearing:

gbE

ddE
tq

D6:67

Seismic displacement demand:


dbE

T2
S T
S T rtg a
2 a
g
4p

D6:68

Equations (D6.67) and (D6.68) and the elastic response spectrum of Eqs (D5.3) yield:
g

if T , TC

gbE 2.5ar
g

if TC  T  TD

gbE 2:5ar
g

(D6.69a)

TC
T

D6:69b

TC TD
T2

D6:69c

if TD , T

gbE 2:5ar

The average seismic shear stress in the bearings:

tbE

P
V
P bE
Ab

D6:70

is obtained as tbE gbEGb:


g

if T , TC

tbE 2.5asb
g

if TC  T  TD

tbE 2:5asb
g

(D6.71a)

TC
T

D6:71b

TC TD
T2

D6:71c

if TD , T

tbE 2:5asb

Figures 6.76.9 show the inuence on the seismic response of the bridge of:
g
g
g

the spectrum corner period TC


the thickness tq in the range 40300 mm
the value of r in the range 212.

Note that, in practice tq seldom exceeds 250 mm and r rarely exceeds 7 or 8, as, around these
values, buckling of the bearings in the seismic design situation may become critical.
p
Figure 6.7 and Eq. (D6.66) show that the period T is proportional to (rtq). In the period range
TC  T  TD , which is of high practical interest, the following are noted:
152

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Figure 6.7. Period T as a function of parameters tq and r of low-damping elastomeric bearings


4.0

= 12

3.5

= 10
=8

Period T: s

3.0

=6

2.5

=4
=3

2.0

=2

1.5
1.0
0.5
0.0
50

100

150
200
Elastomer thickness tq: mm

250

300

Figure 6.8 shows that the same value of the seismic shear strain gb can be obtained with
several pairs of values r and tq below a certain maximum value of r. In fact, gb from
p
Eqs (D6.69b) and (D6.66) is proportional to (r/tq). Note also fromPEqs (D6.64b) and
p
p
(6.64c) that (r/tq) and hence gb  is inversely proportional to ( Abtq) (i.e. to the
square root of the total elastomer volume of the bearings).
Figure 6.9 shows that, among the above pairs of values of r and tq leading to a certain
shear strain demand gb (i.e. pairs giving the same total volume of elastomer), that giving
the most efcient seismic isolation (i.e. the lowest force reduction ratio) is the pair P
with the
Ab).
maximum values of r and tq (i.e. with the maximum value of tq and the minimum

The parts of Figure 6.8 for different values of TC show that the maximum value of r required
for a seismic shear strain gb meeting Eq. (D6.48) drops substantially as TC increases to 0.8 s.
So, for this value of TC and for large design ground accelerations, seismic isolation with lowdamping elastomeric bearings seems unfeasible in practice without supplementary damping.
Even if such a solution proves feasible, its efciency will be very low, as evidenced by
Figure 6.9(c). Possibilities for supplementary damping are discussed in Chapter 7 of this
Designers Guide.
6.10.4.4.3 Practical selection of tq and
Equation (D6.49) may be expanded as

gb,sd gISgbE gbG 0.5gbT  2.0

Ab

(D6.72)

where gbG is the shear strain in the bearing due to shrinkage and creep of the deck and gbT is that
due to temperature variations of and/or through the deck. These strains may be determined as
gbG dG/tq and gbT dT/tq from the relevant displacements dG and dT , which are function of
the maximum distance of a bearing from the centre of stiffness of all the bearings. For
gIS 1.50, Eq. (D6.72) gives

gbE  4/3 (dG 0.5dT)/1.5tq

(D6.73)

On the diagram for the shear strain in Figure 6.8 at the applicable value of TC and for the selected
values of tq , one can mark the points corresponding to tq and the value of gbE required by
Eq. (D6.73). They are on a line approximately parallel to the tq axis, dening pairs of r and tq
satisfying Eq. (D6.72).
Note in the corresponding shear strain diagram that the minimum force (i.e. the optimum
isolation effect) is achieved when r is maximum. This selection maximises the period and the
shear strain in the bearings. The selected pair should, of course, be checked that it meets
153

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Normalised seismic shear strain bE/

Figure 6.8. Seismic shear strain gbE normalised to a Sag/g, as a function of the parameters tq and r of
low-damping elastomeric bearings and of the corner period of the elastic spectrum: (a) TC 0.4 s;
(b) TC 0.6 s; (c) TC 0.8 s
10
9
8
7
6

= 12
= 10
=8
=6

5
4
3
2
1
50

=2 =3

=4

100

150
200
Elastomer thickness tq: mm

250

300

250

300

250

300

Normalised seismic shear strain bE/

(a)
10
9

= 12

= 10
=8
=6

8
7
6
5
4
3
2
1
50

=2 =3
100

=4
150
200
Elastomer thickness tq: mm

Normalised seismic shear strain bE/

(b)
10

= 12
= 10
=8

9
8

=6

7
6
5
4
3

=2 =3

=4

2
1
50

100

150
200
Elastomer thickness tq: mm
(c)

the verications set out in EN 1337-3:2005, notably that of buckling under the fundamental
combination of actions and in the seismic design situation (see Section 6.10.4.2.2), which is
usually more critical than that for the total shear strain in Section 6.10.4.2.3.
The above verications of elastomeric bearings for the seismic design situation need to be carried
out only for the LBDP of the bearings (i.e. for the nominal value of Gb , as it is the critical one for
the deformation-controlled verications of the bearings). However, the maximum forces are the
critical ones for the design of the substructure and for anchoring the bearings, which are both
force-controlled. The corresponding design forces are determined from the UBDP of Gb ,
which is the nominal one increased by 20% if the minimum bearing temperature for seismic
154

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Figure 6.9. Force reduction ratio of the isolation tbE/tbE,max (where tbE,max 2.5asb) as a function of the
parameters tq and r of low-damping elastomeric bearings and of the corner period of the elastic
spectrum: (a) TC 0.4 s; (b) TC 0.6 s; (c) TC 0.8 s

Force reduction ratio bE/bE,max

1.0
0.9
0.8

=2

0.7

=3
=4
=6

0.6
0.5
0.4
0.3
0.2
0.1
0.0
50

= 12 = 10 = 8
100

150
200
Elastomer thickness tq: mm

250

300

250

300

250

300

(a)

Force reduction ratio bE/bE,max

1.0
0.9

=2

0.8

=3
=4
=6

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
50

= 12 = 10 = 8
100

150
200
Elastomer thickness tq: mm
(b)

Force reduction ratio bE/bE,max

1.0
0.9
0.8

=2
=4

=3

=6

0.7
0.6
0.5
0.4
0.3
0.2

= 12 = 10 = 8

0.1
0.0
50

100

150
200
Elastomer thickness tq: mm
(c)

design Tmin,b is not less than 08C, or may be obtained from the appropriate l factor from
Annex JJ of Part 2 of Eurocode 8 if it is.

6.11.

Verication of abutments

6.11.1 General
The abutments and their foundation are designed and veried to stay elastic in the seismic design
situation. Because in these verications the earth pressure action (including seismic effects) is
very important, action effects are normally expressed in the direction of the earth pressure
(i.e. at right angles to the lateral surfaces of the abutment on which the earth pressure acts),

Clause 6.7.1(1) [2]

155

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

which may deviate from the longitudinal or the transverse direction of the bridge. To this end,
peak values of seismic action effects should be available along local axes parallel and at right
angles to the earth pressures. If such local axes have not been introduced at the outset, the
peak values of the effects of the longitudinal and the transverse seismic action are projected
onto the direction of the earth pressures and combined into a single value through Eqs (D5.1)
or (D5.2). The same holds for the seismic action effects at right angles to the earth pressures,
if they are of interest.

Clauses 6.7.2(2),
6.7.2(3), 5.8.2(2),
5.8.2(3) [2]
Clauses 7.3.2.1(1)
7.3.2.1(3), 7.3.2.2(1)
7.3.2.2(7), 7.3.2.3(1)
7.3.2.3(12), 7.3.2.4(1),
7.3.2.4((2), E.3E.9 [3]

6.11.2 Abutments supporting the deck through movable bearings


If the abutment supports the deck on movable bearings, it is not normally included in the global
model for the analysis of the bridge system, even though that model might include piles under the
piers and their interaction with the soil. Such abutments are considered in the global model as
rigid ground, supporting the bearings with the appropriate kinematic restraints or springs.
The analysis gives, then, reactions on/from the bearings, from which the effects of the seismic
action and of the other ones in the seismic design situation on the abutment are derived and
used as loads in independent static analyses and verications of each abutment. The loads/
actions applied to the abutment in this analysis/verication are the following:
1

4
5

156

The non-seismic actions in the seismic design situation:


(vertical) reactions from the bearings on the abutment, derived from the analysis of the
bridge system
gravity loads of the abutment itself and of any earthll over the foundation of the
abutment.
The seismic action effects from the deck on the abutment, as reactions from the bearings
due to the design seismic action. In bridges designed for ductile behaviour, these are
capacity design effects according to Section 6.4.5; in those designed for limited ductile
behaviour, they are the reactions on the bearings from the seismic analysis back-multiplied
by q, to obtain in the end action effects for an effective value of q 1.0.
Pseudo-static inertia forces arising from the mass of the abutment and of the earthll
overlying its foundation. As pointed out in Section 5.4 of this Guide, abutments
supporting the deck on movable bearings are considered as locked in, and their inertia
forces are computed as the pertinent mass multiplied by the design ground acceleration at
the top of the ground at the site, agS (i.e. with q 1.0 and T 0). Note that this is how
seismic action effects due to the mass of the abutment and the earthll over its foundation
should be accounted for, even when the mass and stiffness of these components have been
included in a global analysis model of the bridge.
Any hydrostatic or hydrodynamic pressures, including buoyancy.
The earth pressures (including seismic effects). These pressures are the controlling actions
on the type of abutments considered herein, both under static conditions and in the seismic
design situation. Therefore, the methods for determining earth pressures for free retaining
walls are applicable. According to Annex C of Eurocode 7 (CEN, 2003b), static earth
pressures may be determined on the basis of two limit equilibrium states of the soil:
the passive state, corresponding to the soil resisting movement of the wall against the
earthll
the active one, for the wall giving way to the soil pressure (and possibly to other actions)
and moving away from the earthll.
The pseudo-static MononobeOkabe approach adopted in Part 5 of Eurocode 8 (CEN,
2004c) for the estimation of the seismic effects on earth pressures conveniently uses the
same limit equilibrium states. It is worth bearing in mind for the seismic design situation
the distinction between, on the one hand, movement of the wall relative to the earthll
that denes the passive or active state of earth pressures and the seismic motion of the wall
base and the earthll on the other, which both move with the ground. In fact, the active
earth pressure state of the pseudo-static MononobeOkabe approach develops when the
seismic motion occurs towards the earthll, with the earthll and the wall base by and
large following this motion. In this case the wall body moves away from the earthll under
the inertia forces on the earthll and the wall mass, due to the deformability of the wall
and mainly the rotation of its foundation. Seismic forces from the deck (under point 2
above) are assumed to act in phase with the aforementioned forces, for the most adverse
effect for the abutment, which may fail only when it moves away from the earthll.

Chapter 6. Verication and detailing of bridge components for earthquake resistance

The earth pressures of type 5 acting on a lateral face of the abutment or its wing walls moving
either
(a) against the earthll or
(b) away from it
may be estimated as follows. Those under (a) may be taken as equal to the static earth pressure at
rest, as very large displacements are needed for the passive earth pressure to develop. If the
abutment or the wing wall is free to develop the lateral displacement necessary for an active
state of stress in the backll (see following paragraph), those under (b) are the active earth pressures, including the seismic effects, and are computed with the pseudo-static MononobeOkabe
approach in Part 5 of Eurocode 8. In that calculation, abutments may be considered as nongravity walls, and the effect of the vertical acceleration on earth pressures may be neglected;
the horizontal coefcient kh is taken as equal to the design ground acceleration at the top of
the ground at the site in g: kh agS/g (i.e. the reduction factor r is taken equal to r 1.0).
If, by contrast, the lateral displacement of the abutment or the wing wall is restrained, it is
more realistic and on the safe side to estimate the seismic earth pressures (to be added to the
static ones for the wall at rest) as given in Clause E.9 of Part 5 of Eurocode 8 for rigid structures.
Wing walls monolithic with a large abutment are normally considered to belong in this case. A
signicant lateral displacement, da , is necessary at the top of an abutment for the active state of
stresses to develop in the soil. According to informative Annex C of Eurocode 7, for rotation
about the abutment foundation it takes between 0.4% and 0.5% of the height of the
abutment for loose non-cohesive backll and between 0.1% and 0.2% for dense backll.
About half of these values apply for parallel translation of the abutment. The corresponding
values for the full passive earth pressures in dry soil with an abutment rotating about its
foundation are 725% for loose non-cohesive backll and 510% for dense backll. The
lateral displacement needed for half the full value is 1.54% and 1.12% of the height of the
abutment, respectively (all these values increase by 50100% under the water table); this is
why the static earth pressures at rest are taken to act on the face of an abutment moving
against the earthll. As Part 2 of Eurocode 8 allows consideration of the active earth pressures
on the abutment face moving away from the backll, it is a requirement to check that the
abutment will not fail before the lateral displacement at its top reaches the value, da , where
the MononobeOkabe active earth pressure develops. As the active earth pressure is higher
for lateral displacements less than da , Part 2 of Eurocode 8 requires dimensioning of the
concrete structure of the abutment body and foundation for earth pressures higher than
the active earth pressure; namely, for the value at rest plus 1.3 times the seismic part of the
MononobeOkabe active earth pressure (which is equal to the MononobeOkabe active earth
pressure minus the earth pressure at rest). This increase is not required for the verications of
the foundation soil. The designer should provide, in addition, a clearance between the deck
and the abutment backwall that accommodates a lateral displacement of the top of the
abutment of at least da . Taking into account that the end of the deck is moving away from
the backwall, the required clearance is
dad da  dE dG c2dT

(D6.74)

where dE , dG , c2 and dT assume the same values as in Section 6.8.1.2, and da is the indicative value
quoted above from Part 1 of Eurocode 7 (CEN, 2003b).
Regarding loads of type 5 in the list above, if the face of the abutment towards the bridge is in
contact with water, the hydrostatic pressure minus the hydrodynamic pressure given by normative Annex E in Part 5 of Eurocode 8 under case E.8 should be considered to act on that face,
but with the abutment also moving away from the earthll, which approximately follows the
ground motion as explained above. The minus sign is because the seismic motion is assumed
to be away from the water, which does not move. The case of hydrostatic plus hydrodynamic
pressure on the front side of the abutment, when the abutment and the seismic motion are
against the water, may be ignored alongside the passive earth pressures on the backside of the
abutment. Note that, if the backll itself is saturated and permeable in the time-scale of the
seismic motion (case E.7 in normative Annex E of Part 5 of Eurocode 8), its pore water can
move independently from the ll material. So, the hydrodynamic pressure should be added to
157

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

the earth pressures of the soil. As a result, the hydrostatic plus hydrodynamic water pressure
should be taken to act on the backside of the abutment, while the active earth pressure on it is
proportional to the buoyant unit weight of the soil. Note, however, that only a very coarsegrain ll material may be considered to allow the water to move independently of the ll
during an earthquake.
The static analysis and the verications are carried out considering the components of the seismic
action in two (ideally orthogonal) horizontal directions. Load types 2 and 3 arise from the seismic
action alone, and are taken in these two horizontal directions. Load types 4 and 5 include static
components in xed directions, plus seismic ones which are again taken in the afore-mentioned
horizontal directions. The horizontal forces in load types 2 to 4 due to the seismic action are all
taken with the same sense (positive or negative) along the pertinent horizontal direction. Only if
the abutment and any wing walls are parallel to the transverse and the longitudinal directions of
the bridge, respectively, do the two horizontal directions of the abutment seismic forces coincide
with them. In any other case, the two horizontal directions of the static seismic forces are chosen
ad hoc. For example:
g

Clauses 6.7.3(1)
6.7.3(6), 6.7.3(8) [2]

If the abutment is skewed to the longitudinal axis and does not have wing walls, the static
analyses and the verications are carried out with the components of the seismic action at
right angles and parallel to the abutment.
If the abutment is either at right angle or skewed to the longitudinal axis, possibly
monolithically connected wing walls are usually parallel to the longitudinal direction of the
bridge, along the edges of the roadway. Wing walls of up to 8 m horizontal length are
usually triangular in elevation and do not have their own foundation but cantilever from
the vertical edge of the abutment body. This is quite common for cost reasons, but cannot
be applied to tall abutments. A tall abutment and its wing walls usually form the three
sides of a caisson, with a common foundation behind the abutment body. The main
earthll actions on triangular cantilevering wing walls act at right angles to them, and
amount to the earth pressure at rest combined with that due to a trafc surcharge or the
compaction earth pressure. These actions induce mainly out-of-plane bending of the wing
wall, alongside out-of-plane bending and horizontal tensile forces in the body of the
abutment. Additionally, the wing walls should be veried for the accidental action of
vehicle impact on parapets. The seismic dynamic earth pressure increment, normally
according to case E.9 of Part 5 of Eurocode 8, should be taken as a separate load case
when it acts on each wing wall. Earth pressures on the abutment body are, of course, at
right angles to it. The inuence on the abutment of the seismic earth pressure increment
acting on the skew inside face of the wing wall and obliquely to the abutment may be
estimated approximately. For example, the earth pressure increments estimated for the
abutment may be multiplied by the area of the wall projection on it.
Similar approximations may be used for the caisson-type abutments mentioned above,
for an earthquake at right angles to the abutment. Seismic forces at right angles to the
wing-walls of high and narrow caissons (e.g. with a width-to-height ratio less than 2) may
have to be estimated with the relevant rules for silos in Part 4 of Eurocode 8 (CEN, 2006).

6.11.3 Abutments rigidly connected to the deck


An important case here is that of a deck monolithic with the abutments (integral). The other
case is that of an abutment connected to the deck via xed bearings or seismic links (including
shear keys) designed to carry the seismic action, in lieu of the bearings. For the connection to
be considered as rigid (and for clause 6.7.3 of Part 2 of Eurocode 8 to apply), it should be so
in both horizontal directions. The length and span limitations set out in Section 4.5.3 of this
Designers Guide for integral bridges also hold for decks connected to the abutments through
xed bearings or seismic links designed to carry the seismic action.
Most of the rules in clause 6.7.3 of Part 2 of Eurocode 8 (namely the q value of 1.50, the earthll
reaction to be considered according to the next paragraph and the upper limit to the total
design seismic displacement in the last paragraph of this section) refer to short integral
bridges. Where the connection of the deck to the abutment is horizontally movable or
exible, at least in the longitudinal direction (as is always the case in longer bridges), none
of these rules apply.

158

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Regarding the activation of earthll reaction against the backside of abutments rigidly connected
to the deck, note the following:
g

In the longitudinal direction, reaction is activated on the face of the backwall towards the
backll. Part 2 of Eurocode 8 recommends using over it equivalent springs with a stiffness
corresponding to the actual geotechnical conditions. To cover uncertainty in the
distribution of the seismic force among any piers and the reacting abutment, it also
recommends using upper- and lower-bound soil compliance properties. Lower-bound
earthll stiffness leads to the maximum seismic forces for the piers and their foundations
and to maximum seismic displacements of the abutment which are subject to the damage
limitation check at the end of the present section. Upper-bound stiffness may be critical
only for the abutment in bending and the adjacent region of the deck.
In the transverse direction, the area where earthll reaction may be activated is the internal
face of wing walls that are monolithically connected to the abutment and move towards
the earthll. Again, upper- and lower-bound soil compliance estimates should be used. The
lower-bound stiffness of the soil may be close to zero if the ratio width/height of the
earthll is low in the transverse direction. Active dynamic earth pressures should anyway
be considered on the opposite wing wall (see Section 6.11.2).

A global model is normally used for the analysis, with a fairly detailed discretisation of the deck
and the abutments. Springs model the soil against which the abutments and any wing walls may
push, with use of upper- and lower-bound estimates of the soil stiffness according to the paragraph above. Bridges with the deck exibly supported on the abutments in the longitudinal direction, but restrained there transversely by seismic links (shear keys), should be analysed for
transverse seismic action with a global model involving soil springs at the interior surface of
the wing walls. In view of the need to consider upper- and lower-bound estimates of soil stiffness,
this is quite taxing. However, there is no need for such a model if the abutments do not have
(signicant) wing walls pushing against the backll.
Bridges with the deck monolithic with the abutments often meet the criteria for the application of
the linear static analysis with the fundamental mode method. With inertia forces on the masses of
the structure estimated this way, the soilstructure system is then analysed for the simultaneous
action of gravity loads on these masses and of earth pressures (static at rest, plus seismic where
pertinent) and hydrostatic and hydrodynamic ones, where applicable. The seismic earth pressures
and the hydrodynamic pressures are computed and applied where pertinent according to Section
6.11.2, but always in the same horizontal direction and with the same sense of action as the inertia
forces on the masses of the structure.
To limit damage to the backll behind an abutment rigidly connected to the deck, Part 2 of
Eurocode 8 requires checking that the displacement of the abutment towards the backll,
computed from the analysis results according to Eq. (D5.68), does not exceed a certain limit.
This limit is a Nationally Determined Parameter, with a recommended value of 60 mm in
bridges of importance class II or just 30 mm for those of importance class III. No limit is set
for bridges of importance class I.

6.12.

Verication of the foundation

6.12.1 Design action effects


For the purpose of the verication of resistance, the design action effects on the foundation
should be determined as follows (see Sections 6.3.2 and 6.7.2):
g

Clause 6.7.3(7) [2]

Clauses 4.2.4.4(2)e,
5.8.2(2)5.8.2(4) [2]

For bridges designed for limited ductile behaviour (q  1.5) or for those with seismic
isolation, the design action effects are those from the analysis for the seismic design
situation, Eq. (D6.1), with the seismic action effects computed for elastic response (i.e.
with a behaviour factor of 1.0).
For bridges designed for ductile behaviour (q . 1.5):
if the linear analysis with the design response spectrum is applied, the capacity design
procedure in Section 6.4.2 should be applied, for exural plastic hinges developing in the
piers
159

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

if nonlinear analysis is used, the design action effects are those from the analysis, but the
design resistance (computed with material partial factors, gM) is divided by the
Nationally Determined Parameter safety factor, gBd1 , with a recommended value of
gBd1 1.25 (see Sections 6.7.2 and 6.9.2).

Clause 5.4.1 [3]

6.12.2 Shallow foundations


6.12.2.1 Stability verications
The verications of shallow foundations concern sliding and the bearing capacity of the soil in
the seismic design situation.
For the verication against sliding, the total design horizontal force should satisfy the following:
VEd  FH1 FH2 0:3FB

D6:75

where FH1 is the friction over the base of the footing, equal to NEd tan(d)/gM; FH2 is the friction
over the lateral sides of an embedded foundation; FB is the ultimate passive resistance of an
embedded foundation; NEd is the vertical design force acting on the foundation; d is the
friction angle between the foundation and the soil; and gM is the partial factor, taken equal to
gw, with a recommended value of 1.25.
Note that, although full friction on the base and the lateral sides of the foundation may be
considered to be mobilised, relying on more than 30% of the total passive resistance is not
allowed. The rationale for this limitation is that, for mobilisation of the full passive resistance,
a displacement of signicant magnitude is necessary to take place that does not comply with
the usual performance goals set forth for bridges. The values of FH2 and FB should correspond
to the lowest top level of the ground surface expected in the seismic design situation.
Note that, under certain circumstances, sliding may be acceptable, as an effective means for
dissipating energy and shielding the superstructure by limiting the forces that enter it (as in
base isolation systems). Furthermore, a numerical simulation generally shows that the amount
of sliding is limited. For this situation to be acceptable, the ground characteristics should
remain unaltered during the seismic excitation. Additionally, sliding may not affect the functionality of the bridge. Since soils under the water table may be prone to pore pressure build-up,
which affects their shear strength, sliding is only tolerated when the foundation is located
above the water table. It should be further pointed out that the predicted foundation displacement when sliding is allowed strongly depends on the friction coefcient between the footing
surface and the soil, which in turn depends on the surface material, its drainage conditions
and the construction method. If reliable estimates are necessary, in-situ tests are warranted.
Although not required by Eurocode 8, it is common practice to keep the resultant force at the
level of the soilfooting interface within one-sixth to one-third of the footing width from the
centre; if a linear distribution of soil stresses is assumed under the footing, these two limits
are equivalent to not allowing uplift, or restricting it to one-half of the foundation, respectively.
Nowadays, uplift is more commonly tolerated, because it is recognised that rocking of the
foundation reduces the seismic forces that enter the structure and therefore protects it.
However, to avoid yielding of the soil under the loaded edge and the permanent settlement
and tilt of the foundation it entails, rocking must be restricted to very good soil conditions.
To evaluate this behaviour for a spread footing, it is recommended that nonlinear static
(pushover) analysis is conducted to establish its momentrotation characteristics, including the
effects of uplift of the footing and soil yielding. Analytical results from such analyses not only
provide rotational stiffness parameters but also depict the geotechnical mode of ultimate
moment capacity. For spread footings, uplift is the most severe form of nonlinearity, albeit
tolerated by Eurocode 8. The foundation cannot develop overturning moments higher than
the ultimate moment capacity and its momentrotation curve becomes distinctly nonlinear
well before the ultimate moment (see Figure 5.20 for examples).
Clause 5.4.1.1,
Annex F [3]

160

In addition to the sliding verications, the bearing capacity of a foundation under seismic
conditions should be checked, considering the inclination and eccentricity of the force acting
on the foundation, as well as the effect of the inertia forces developed in the soil medium by

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Figure 6.10. Ultimate load surface for the foundation bearing capacity

the passage of the seismic waves. A general expression, Eq. (D6.76), is provided in Annex F of
Part 5 of Eurocode 8, derived from theoretical limit analyses of a strip footing (Pecker, 1997).
The condition for the foundation to be safe against bearing capacity failure simply expresses
that the design forces NEd (vertical), VEd (horizontal), MEd (overturning moment) and the soil
seismic forces should lie within the surface depicted in Figure 6.10 and expressed by

c  c

 c0 
c
1  eF T bV T
1  f F M gM M

b

d  1
0
0
 a 
 c 
k k
k k
N
1  mF
N
N
1  mF
N

D6:76

with the following denitions:


N
g

gRd VEd
Nmax

gRd MEd
BNmax

D6:77

for purely cohesive or saturated cohesionless soils


F

gRd NEd
Nmax

gRd rag SB
Cu

D6:78a

for purely dry or saturated cohesionless soils without signicant pore pressure build-up
F

gRd ag S
g tan f0d

D6:78b

where Nmax is the ultimate bearing capacity of the foundation under a vertical concentric force, as
estimated by any reliable method (strength parameters, empirical correlations from eld tests,
etc.); B is the foundation width; F is the dimensionless soil inertia force; gRD is the model
factor; r is the unit mass of the soil; ag is the design ground acceleration on type A ground,
Eq. (D3.3) in Section 3.1.2.2; S is the soil factor, see Table 3.3 in Section 3.1.2.3; Cu is the soil
undrained shear strength, cu , for a cohesive soil or the cyclic undrained shear strength, tcy,u ,
for a saturated cohesionless (including the material partial factor, gM); and fd0 is the design
angle of the shearing resistance of cohesionless soil (including the partial factor, gM).
The coefcients in lower case in Eq. (D6.76) (a, b, k, cM and cT) are numerical values that depend
on the soil type according to Table 6.2 (from Part 5 of Eurocode 8). The model factor gRD
161

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 6.2. Values of numerical parameters in Eq. (D6.76) (Part 5 of Eurocode 8)

a
b
c
d
e
f
m
k
k0
cT
cM
0
cM
b
g

Purely cohesive soil

Purely cohesionless soil

0.70
1.29
2.14
1.81
0.21
0.44
0.21
1.22
1.00
2.00
2.00
1.00
2.57
1.85

0.92
1.25
0.92
1.25
0.41
0.32
0.96
1.00
0.39
1.14
1.01
1.01
2.90
2.80

reects the uncertainties in the theoretical model for the seismic verication of the bearing
capacity, and as such it should be larger than 1.0; it is also meant to recognise that certain
permanent foundation displacements may be tolerated (in which case Eq. (D6.76) is violated),
and then it may be less than 1.0. Tentative values that intend to combine both effects are
proposed in Annex F in Part 5 of Eurocode 8 and repeated in Table 6.3, reecting that for the
most sensitive soils (loose saturated soils) the model factor should be higher than for stable
ones (medium-dense sand).
More recent studies have shown that Eqs (D6.76)(D6.78), as well as these tabulated values, hold
also for circular footings, provided that the ultimate vertical force under vertical concentric
force, Nmax entering Eq. (D6.77), is computed for a circular footing and that the footing width
is replaced by the footing diameter (Chatzigogos et al., 2007). Although Eq. (D6.76) does not
look familiar to geotechnical engineers who are more accustomed to the classical bearing
capacity formula with correction factors for load inclination and eccentricity, it reects the
same aspect of foundation behaviour. This verication is similar to the use of interaction
diagrams in structural design for cross-sections under combined axial force and bending
moment.
6.12.2.2 Structural design of the footing
A footing is usually designed by treating it as a cantilever supported at the verication
sections, or as a simple beam, or as a continuous beam between the columns of rigid frame
piers. However, as footings may behave as a slab with an internal redistribution of stresses,
they may be considered as two-way beams for design purposes. The footing should have
sufcient thickness to be regarded as rigid relative to the underlying soil. Then, a linear
distribution of soil stresses or a distribution corresponding to an appropriate no-tension
Winkler spring model may be assumed for the determination of the internal forces of the
footing at its structural ULS.
As the footing is designed at the ULS as a beam, unless a nite element slab model is used for
the analysis, it is necessary to dene an effective width of the footing for the ULS in bending;

Table 6.3. Model factors for Eqs (D6.77) and (D6.78) (Part 5 of Eurocode 8)

162

Medium-dense to dense
sand

Loose dry
sand

Loose saturated
sand

Non-sensitive
clay

Sensitive
clay

1.00

1.15

1.50

1.00

1.15

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Table 6.4. Effective width of footings according to JRA (2002)


Effective footing width
Bottom reinforcement of footing
Top reinforcement of footing

bB
b hc 1.5d  B

(JRA, 2002) provides the estimates of effective width in Table 6.4 as a function of the pier width,
hc , and the effective footing depth, d.
6.12.3 Pile foundations
6.12.3.1 Pile internal forces
Piles and piers should be veried for the effects of the inertia forces transmitted from the superstructure to the pile heads and also for the effects of kinematic forces due to earthquake-induced
soil deformations. However, according to Part 2 of Eurocode 8, kinematic interaction needs to be
considered only when the pile crosses consecutive layers of sharply contrasting stiffness.
According to Part 5 of Eurocode 8, consideration of kinematic forces in piles is required when
the weak layer consists of soft deposits (as in ground types D, S1 or S2) with layers of sharp
stiffness contrast, the design acceleration exceeds 0.10g and the bridge is of importance above
ordinary.

Clauses 5.4.2(1)
5.4.2(6) [3]
Clause 6.4.2.2(2)c [2]

For long exible piles, the effects of kinematic interaction may be evaluated assuming that the
piles follow the ground displacement, which may be calculated either from a 1D site response
analysis with a computer code such as SHAKE (Idriss and Sun, 1992) or, in a homogeneous
soil layer overlying bedrock, from a simplied analytical formulation with the soil layer
assumed to respond in its fundamental mode of vibration and the pile modelled with the
appropriate boundary conditions at its tip and head (hinged, pinned). The displacement of the
fundamental mode of vibration of the soil layer is given by
d z d sin

 pz
2H

D6:79

where H is the layer thickness and d is the relative ground displacement between the top and the
base of the soil layer. This relative ground displacement can be read from the rock response
spectrum (see Section 3.1.2.3 of this Guide) at the fundamental frequency of the soil layer
( f vs/4H) for the applicable shear wave velocity, vs , and damping ratio (see Table 3.5).
The effects of the inertia forces are obtained from the dynamic analysis. If the piles are modelled
as beams on a Winkler foundation, these results are directly obtained from the analysis. If the
soilstructure interaction is modelled as linear, it should, however, be checked that the
pressure applied by the piles to the soil does not exceed the pertinent ultimate capacity. If it
does, the analysis should be redone with weaker springs wherever the ultimate soil bearing
capacity is exceeded, until compatibility between computed displacements and the pressure is
achieved. If the pile foundation is modelled using the concept of the impedance matrix
introduced in Section 5.5.1.6 of this Guide, the internal forces in the piles can be obtained
only through a separate model for the pile, subjected to the same limitations as above.
The values of the action effects arising from inertial and kinematic loading may be combined
using the SRSS (square root of the sum of the squares) rule, unless the fundamental periods
of the soil layer and the bridge are close to each other. If they are, the sum of their absolute
values is more appropriate.
6.12.3.2 Structural design
Piles are generally designed to remain elastic: a q factor of 1.0 in bridges designed for limited
ductile behaviour; and through capacity design in those designed for ductility. However, with
the large bending moment that develops at the connection of a pile to its cap, designing the
pile to remain elastic there may not be feasible. It is therefore more economical and often
safer to design that location to develop a plastic hinge. This is indeed allowed in Part 2 of

Clauses 4.1.6(7),
6.4.2(1)6.4.2(4) [2]
Clause 5.4.2(7) [3]

163

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 6.11. Connement of the compression zone by the soil pressure for in-ground hinges

Soil
pressure

Moment

Cracking and soil pressure

Eurocode 8, with the q factors of rows 3 and 4 in Table 5.1. In this case, locations of potential
plastic hinges should be detailed:
(a) the top of the pile to a distance of three pile diameters, D, from the underside of a pile cap
that is restrained against rotation in a vertical plane
(b) a length 2D of the pile on each side of the location of maximum bending moment (as
estimated taking into account the effective pile exural stiffness, the lateral soil stiffness
and the rotational stiffness of the pile group at the pile cap) and of an interface between
soil layers with a marked stiffness contrast.
Such detailing consists of conning reinforcement with a mechanical reinforcement ratio as
specied in Table 6.1 for piers of ductile behaviour. At locations of type (b) above, Eurocode 8
does not reduce the required connement steel to account for the experimentally found (Budek
et al., 1997) benecial effect of the soil pressures exerted on the compression side of the pile
(Figure 6.11). In such locations it requires a pile vertical reinforcement of not less than that
placed at the pile head.
Piles where plastic hinges are allowed should be veried for capacity design shear forces. If the
bridge is designed using nonlinear analysis, the head of the pile should also be veried as a
plastic hinge according to Section 6.7.2. In such a verication, the shear span, Ls , of the pile
may be taken as the distance from the point of contraexure to the pile cap.

6.13.
Clause 4.1.4(12) [3]

Liquefaction and lateral spreading

6.13.1 Introduction
Most seismic codes and current practice are against placing foundations on liqueable deposits.
According to Part 5 of Eurocode 8:
if soils are found to be susceptible to liquefaction and the ensuing effects are deemed capable of
affecting the load bearing capacity or the stability of the foundations, measures . . . shall be
taken to ensure foundation stability.
When designing a foundation on a potentially liqueable deposit three situations need to be
analysed:
g

164

Stage 1: during which limited pore pressures have developed and the soil retains its
original stiffness and strength; at this stage the forces applied to the foundation are most
likely less than the maximum forces that would develop during the earthquake if
liquefaction has not occurred; the response of the bridgefoundation system is purely
dynamic.

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Stage 2: here, considerable excess pore pressures have developed (typically Du/sv0 . 0.5,
where sv0 is the vertical effective overburden) but no widespread liquefaction has yet
occurred; at this stage the soil stiffness is drastically reduced but the soil still preserves a
non-zero shear stiffness, allowing the seismic waves to propagate through the soil and
affect the foundation; the response of the bridgefoundation system is still predominantly
dynamic, but signicant nonlinearities take place that reduce the input motion and
therefore the dynamic forces.
Stage 3: this takes place towards the end or after the earthquake shaking has ended; full
liquefaction has developed, and lateral spreading of the liqueed soil may take place under
certain conditions, such as an inclination of the soil layers or of the ground surface,
resulting in a quasi-static, gravity-induced loading to the foundation. The response is
essentially static.

6.13.2 Shallow foundations


For shallow (footing or mat) foundations, the upward ow of water towards the ground surface
induces a signicant, if not a total, loss of soil resistance. Because the bearing layers are at the
ground surface, the strength reduction occurring in stage 2 or 3 causes loss of bearing
capacity, accompanied not only by vertical settlement but also, in some instances, signicant
tilt. The foundation movements that accompany full development of pore pressures are
unpredictable. Therefore, countermeasures need to be implemented. There is no alternative to
improving the ground conditions.
6.13.3 Pile foundations
6.13.3.1 Development of design approaches
For pile foundations the situation is somewhat different, as the load-resisting elements can be
embedded below the potentially liqueable strata. Therefore, it is feasible to design pile foundations to accommodate soil liquefaction. This has been done in practice at least to accommodate stages 1 and 2. However, until the Kobe earthquake of 1995, it was not foreseen to
design pile foundations to accommodate lateral spreading (stage 3), and signicant soil improvement needed to be implemented to protect the foundations of bridge piers. Soil improvement may
be expensive, especially when a wide area is affected by lateral spreading. For instance, the north
approach viaduct of the RionAntirrion Bridge in Greece is located in a zone where extensive
lateral spreading was predicted under the design earthquake; a costbenet analysis showed
that designing the piles to resist the forces induced by the soil displacement was more costeffective than improving the ground conditions. It is only recently that this type of analysis
Figure 6.12. Pile conguration due to liquefaction
Free field soil
deflection

Non-liquefied
layer

Liquefied
layer

Deformed
shape of pile

Non-liquefied
layer

165

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 6.13. Idealisation of ground ow for the seismic design of bridge foundations

Non-liquefiable
layer

qNL

HNL

Liquefiable
layer

qL
HL

Non-liquefiable layer

has been possible. Extensive back analyses of damage to pile foundations during the Kobe earthquake resulted in rational design methodologies (Finn, 2005). The situation to be analysed is
depicted in Figure 6.12. After liquefaction, if the residual strength of the soil is less than the
static shear stress caused by a sloping site or a free surface, such as a river bank, signicant
down-slope movement may occur. The moving soil exerts damaging pressures against the
piles, especially if a non-liqueed layer rides on top of the liqueed layer. Basically, there are
two different approaches to evaluate the forces: a force-based approach and a displacementbased approach.
6.13.3.2 Force-based approach
A force-based analysis is recommended in JRA (2002) for pile foundations in liqueed soils:
g

the non-liqueed surface layer is assumed to apply a passive pressure, qNL , on the
foundation

Figure 6.14. Winkler spring model for lateral spreading analysis

Ground
displacement

Liquefied
layer

166

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Table 6.5. Reduction coefcients for soil springs due to liquefaction according to JRA (2002)
Safety factor, FL

Depth from ground


surface, x: m

Dynamic shear strength ratio, R cwRL


R  0.3

R . 0.3

FL  1/3

0  x  10
10 , x  20

0
1/3

1/3
1/3

1/3 , FL  2/3

0  x  10
10 , x  20

1/3
2/3

2/3
2/3

2/3 , FL  1

0  x  10
10 , x  20

1/3
1

1
1

the liqueed layer is taken to apply a pressure, qL , less than the equivalent hydrostatic
pressure.

Figure 6.13 illustrates the forces acting on the foundation. It has been found that the pressure
exerted by the liqueed layer may be taken as equal to 30% of the overburden pressure. These
ndings have been conrmed by centrifuge tests (Dobry and Abdoun, 2001), which, in
addition, showed that the moments in the pile are dominated by the lateral pressure from the
non-liqueed layer. Therefore, to minimise the bending moments in the piles arising from the
pressures exerted by the top non-liqueed soils, it is highly desirable to locate the pile cap
above the ground surface, without contact with the in-place soils.
6.13.3.3 Displacement-based approach
Following this approach, forces are not applied to the piles. Instead, free eld displacements are
imposed at the free ends of the springs in the Winkler model in Figure 6.14. The method requires
knowledge of the free eld displacements, which may be estimated via predictor equations
described in Section 3.4.5 of this Guide.
The displacement-based approach relies on two input data: the predictor equations giving the
amplitude of lateral spreading and the values of the spring stiffness to be used in the pile
model. In Japanese practice, the reduction in the spring stiffness for use in liqueable soils
depends on the factor of safety against liquefaction, FL . The recommended reduction factors
are given in Table 6.5 as a function of the product of the resistance to liquefaction, RL , and of
parameter cw , which for the reference seismic action dened in Eurocode 8 can be taken as
cw 1.0

for RL  0.1,

(D6.80a)

cw 3.3RL 0.67

for 0.1 , RL  0.4

(D6.80b)

cw 2.0

for 0.4 , RL

(D6.80c)

There is no commonly accepted practice in North America for the appropriate modelling of
degraded spring stiffness. The basis of most analysis is a degraded form of the American
Petroleum Institute py curves (API, 1993). The practice is to multiply the py curves by a
uniform degradation coefcient b, which ranges from 0.3 to 0.1 (Finn, 2005).
There is a great uncertainty in the denition of the free eld displacements used as input data to
the analysis. The predictor equations (Youd et al., 2002) are strongly empirical, and based on few
observations. Consequently, the force-based approach is often preferred and recommended
(JRA, 2002).
REFERENCES

API (1993) Recommended Practice for Planning, Designing and Constructing Fixed Offshore
Platforms. American Petroleum Institute, Washington, DC.
Biskinis DE and Fardis MN (2010) Flexure-controlled ultimate deformations of members with
continuous or lap-spliced bars. Structural Concrete 11(2): 93108.
167

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Biskinis DE and Fardis MN (2012) Effective stiffness and cyclic ultimate deformation of circular
RC columns including effects of lap-splicing and FRP wrapping. 15th World Conference on
Earthquake Engineering, Lisbon.
Budek AM, Benzoni G and Priestley MJN (1997) Experimental Investigation of Ductility of InGround Hinges in Solid and Hollow Precast Prestressed Piles. Division of Structural Engineering,
University of California, San Diego, CA. Report SSRP-97.
CEN (Comite Europeen de Normalisation) (2000) EN 1337-2:2000: Structural bearings Part 2:
Sliding elements. CEN, Brussels.
CEN (2002) EN 1990:2002: Eurocode: Basis of structural design (including Annex A2: Application
to bridges). CEN, Brussels.
CEN (2003a) EN 1991-1-5:2003: Eurocode 1: Actions on structures Part 15: General actions
Thermal actions. CEN, Brussels.
CEN (2003b) EN 1997-1:2003: Eurocode 7: Geotechnical design Part 1: General rules. CEN,
Brussels.
CEN (2004a) EN 1998-1:2004: Eurocode 8 Design of structures for earthquake resistance Part
1: General rules, seismic actions and rules for buildings. CEN, Brussels.
CEN (2004b) EN 1992-1-1:2004: Eurocode 2: Design of concrete structure. Part 1: General rules
and rules for buildings. CEN, Brussels.
CEN (2004c) EN 1998-5:2004: Eurocode 8 Design of structures for earthquake resistance Part
5: Foundations, retaining structures, geotechnical aspects. CEN, Brussels.
CEN (2005a) EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance Part
2: Bridges. CEN, Brussels.
CEN (2005b) EN 1992-2:2005: Eurocode 2: Design of concrete structure. Part 2: Bridges. CEN,
Brussels.
CEN (2005c) EN 1998-3:2005: Eurocode 8 Design of structures for earthquake resistance Part
3: Assessment and retrotting of buildings. CEN, Brussels.
CEN (2005d) EN 1337-3:2005: Structural bearings Part 3: Elastomeric bearings. CEN, Brussels.
CEN (2006) EN 1998-4:2006: Eurocode 8: Design of structures for earthquake resistance Part 4:
Silos, tanks and pipelines. CEN, Brussels.
CEN (2009) EN 15129:2009: Antiseismic devices. CEN, Brussels.
Chatzigogos CT, Pecker A and Salencon J (2007) Seismic bearing capacity of circular footing on an
heterogeneous cohesive soil. Soils and Foundations 47(4): 783797.
Constantinou MC, Kalpakidis I, Filiatrault A and Ecker Lay RA (2011) LRFD-based Analysis and
Design Procedures for Bridge Bearings and Seismic Isolators. Department of Civil, Structural and
Environmental Engineering, State University of New York, Buffalo, NY. Technical Report
MCEER-11-004:2011.
Dobry R and Abdoun T (2001) Recent studies of centrifuge modeling of liquefaction and its effect on
deep foundations. Proceedings of the 4th International Conference on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics (Prakash S (ed.)), San Diego, CA, pp. 2631.
Elwi AA and Murray DW (1979) A 3D hypoelastic concrete constitutive relationship. Journal of
Engineering Mechanics Division of the ASCE 105(EM4): 623641.
Fardis MN (2009) Seismic Design, Assessment and Retrotting of Concrete Buildings (Based on ENEurocode 8). Springer-Verlag, Dordrecht.
Finn WDL (2005) A study of piles during earthquakes: issues of design and analysis. The tenth
Mallet Milne Lecture. Bulletin of Earthquake Engineering 3(2): 141234.
Gupta AK and Singh MP (1977) Design of column sections subjected to three components of
earthquake. Nuclear Engineering and Design 41: 129133.
Idriss IM and Sun JI (1992) SHAKE 91: A Computer Program for Conducting Equivalent Linear
Seismic Response Analyses of Horizontally Layered soil Deposits. Program Modied Based
on the Original SHAKE Program Published in December 1972 by Schnabel, Lysmer and Seed.
Center of Geotechnical Modeling, Department of Civil Engineering, University of California,
Davis, CA.
JRA (2002) Design Specications for Highway Bridges, Part V. Seismic Design. Japanese Road
Association, Tokyo.
Katsaras CP, Panagiotakos TB and Kolias B (2009) Effect of torsional stiffness of prestressed
concrete box girders and uplift of abutment bearings on seismic performance of bridges. Bulletin
of Earthquake Engineering 7(2): 363376.
Mander JB, Priestley MJN and Park R (1988) Theoretical stressstrain model for conned
concrete. ASCE Journal of Structural Engineering 114(8): 18041826.
168

Chapter 6. Verication and detailing of bridge components for earthquake resistance

Paulay T and Priestley MJN (1992) Seismic Design of Reinforced Concrete and Masonry Buildings.
Wiley, New York.
Pecker A (1997) Analytical formulae for the seismic bearing capacity of shallow strip foundations.
In Seismic Behavior of Ground and Geotechnical Structures (Seco e Pinto PS (ed.)). Balkema,
Rotterdam.
Richart FE, Brandtzaeg A and Brown RL (1928) A Study of the Failure of Concrete Under
Combined Compressive Stresses. University of Illinois Engineering Experimental Station,
Champaign, IL. Bulletin 185.
Stanton JF, Roeder CW, Mackenzie-Helnwein P et al. (2008) Rotation Limits for Elastomeric
Bearings. Transportation Research Board, National Research Council, Washington, DC.
NCHRP Report 596.
Youd TL, Hansen CM and Bartlett SF (2002) Revised multilinear regression equations for
prediction of lateral spread displacement. Journal of Geotechnical and Geoenvironmental
Engineering 128(12): 10071017.

169

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.171

Chapter 7

Bridges with seismic isolation


7.1.

Introduction

The concept of seismic isolation is more familiar to designers of bridges than of buildings or other
structures. Indeed, to avoid restraining thermal elongation or shortening and concrete shrinkage,
bridge designers have since long adopted the practice of separating the deck from the abutments
and some piers, by introducing longitudinally movable bearings (rollers, sliders or even elastomeric bearings). From that point it takes only a short leap to a continuous isolating interface
between the deck and all of its supporting elements, in order to reduce the transmission of
seismic motion from the ground to the deck.

7.2.

Objective, means, performance requirements and conceptual


design

7.2.1
Objective and means
As pointed out above, the aim of seismic isolation is to reduce the seismic inertia forces on the
main mass of the bridge. To this end, an interface of isolating devices, termed the isolation interface, is introduced between the superstructure (normally the deck) and the substructure (i.e. the
abutments and piers). Part 2 of Eurocode 8 (CEN, 2005a) uses the generic term isolators for
the units of the isolation system. Isolators may offer one or more of the following capabilities:

Clause 7.1(1) [2]

Vertical supporting function with high stiffness, to ensure the safe transfer of the vertical
reactions without deck deformation.
2 Transfer of horizontal forces with the effective displacement stiffness much less than that
of the corresponding element of the substructure (pier or abutment). These horizontal
forces act usually but not always as elastic restoring forces for the isolation system.
3 Enhanced dissipation of seismic input energy through increased damping.
1

The reduction of the horizontal stiffness of the total system Keff (point 2 above) increases its
p
fundamental period Teff 2p (M/Keff) relative to that of the initial structure with horizontally
xed connections, Tf . The acceleration spectrum in Figure 7.1(a) shows that this period shift
brings about a substantial reduction in the spectral acceleration, especially when Teff  TC ,
where TC is the corner period between the constant pseudo-acceleration and the constant
pseudo-velocity ranges of the elastic spectrum of Eqs (D3.8). However, the displacement
spectrum in the same gure (derived from Eq. (D3.1)) clearly shows that the displacement is
also increased substantially.
An increase in damping to a value jeff above the default value of 5% of the elastic spectrum,
owing to the additional damping offered by the isolation system, reduces both the displacements,
p
as shown in Figure 7.1(b), and the forces, by the damping modication factor h [10/(5 jeff)]
(used also in Eqs (D3.8) and (D5.48), but allowed to be as low as 0.40 for seismic isolation). The
combination of increased damping and a period shift is a very efcient means to a substantial
reduction in seismic forces without an inconvenient increase in displacements.
7.2.2
Performance requirements
7.2.2.1 General requirements
Bridges with seismic isolation must meet the two general performance requirements set by Part 2 Clauses 2.2, 7.3(1) [2]
of Eurocode 8 for all bridges: the non-collapse and the limitation of damage requirements (see
Section 2.2 of this Designers Guide). However, specically for bridges with seismic isolation,
additional requirements are set, as detailed in the following sections.
171

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.1. Means for achieving seismic isolation: (a) period shift; (b) increased energy dissipation
(damping)
a

Period shift
d

a
dDisplacement spectrum

ag

d=a
To Tc

Teff
(a)

Clauses 7.3(4),
7.6.2(1), 7.6.2(2) [2]

o to eff

TD

( )
T

Teff
(b)

7.2.2.2 Increased reliability of the isolation system


All components of the isolation system must have increased reliability relative to the other components of a bridge, both in terms of displacement capability and of force resistance. The enhanced
reliability is deemed to be available if the displacement capacity of the isolators can accommodate
a total design displacement in the seismic design situation, dEd , from Eq. (D6.36) in Section 6.8.1.2,
computed with the design seismic displacement, dE , from Eq. (D5.48) in Section 5.9, multiplied by
a safety factor gIS having a recommended value of 1.50. The resistance of the isolators and their
anchorage should accommodate the force developed at this increased displacement level.
The rationale behind the increased reliability demanded from the components of the isolation
system can be seen by comparing the design parameters dening the corresponding ultimate
limit states (ULS):
g

Clauses 7.3(2),
7.6.3 [2]

The ULS of ductile or limited ductile components is dened by resistances equal to the
characteristic design values (i.e. lower 5% fractiles) divided by the material safety factors
(with recommended values of 1.50 for concrete and 1.15 for steel). Therefore, these
components possess substantial safety margins. Should they momentarily and locally
exceed their force limits, the effects will normally be reversible and not catastrophic, owing
to their inherent ductility margins.
The components of the isolation system do not enjoy the above favourable situation. Their
ULS is practically dened by the geometric condition of the displacement capacity,
without margins. Exceeding this capacity is irreversible and usually catastrophic for the
isolator and probably for the bridge as well.

7.2.2.3 Behaviour of the substructure


As pointed out in Section 7.2.1, the main benet of the isolation system is the high horizontal
exibility at the connection between the deck (which accounts for most of the system mass)
and its supporting elements (piers and abutments). This produces a fundamental natural mode
of vibration in the isolated horizontal direction with a substantially longer period than higher
modes. The response in the isolated mode(s) dominates the seismic response of the bridge. In
addition, the damping introduced by the isolation system acts directly on the displacements
and forces of the dominant mode(s) and is, therefore, quite efcient for dissipating energy and
reducing the global seismic response. Consequently, the seismic response of the substructure is
substantially reduced, and may be undertaken via elastic or limited ductile behaviour.
One might consider attempting a further reduction in the force response by providing for a
ductile behaviour of the substructure, allowing plastic hinges to form there and exploiting
their inelastic behaviour (partial isolation, instead of full). However, this, in general, will
not be benecial to the bridge for the following reasons, due to which Eurocode 8 does not
allow partial isolation neither for buildings (Part 1 (CEN, 2004)) nor for bridges (Part 2):
g

172

Partial isolation simply transfers part of the displacement demand from the isolation
system to the substructure, to be accommodated there by inelastic deformation (i.e.

Chapter 7. Bridges with seismic isolation

damage) of certain elements. In addition to the damage, inelastic deformation entails


larger uncertainty than displacements within the (increased if necessary) capacity of the
isolation system.
The inelastic forcedisplacement relation of the entire system no longer has the simple
linear shape of a fully isolated structure, where the dominant linear branch depends
mainly on the post-elastic stiffness of the isolation system and to a minor degree on the
elastic stiffness of the piers. This branch starts from the early yielding of the isolation
system and runs straight to its displacement capacity. By contrast, in partial isolation a
new linear branch starts at the yield displacement of each pier, with the last one
continuing until the minimum ultimate displacement capacity among all piers or the rst
displacement capacity is reached among the isolators (assumed here without restraining
devices). The determination of such a complex forcedisplacement law requires fairly
accurate evaluation of many parameters, and would in any case be subject to large
uncertainties.
Transferring signicant displacement demands from the isolation system to plastic hinges
in the substructure reduces signicantly the benecial effect of the energy dissipated by the
isolators, which is approximately proportional to the square of their displacement.
An important reason for choosing to seismically isolate a structure is to drastically
minimise or practically eliminate damage in the seismic design situation, by keeping the
structure within its elastic limits. This advantage is lost for partial isolation.

7.2.2.4 Consideration of the variability of the design properties of the isolation


system
The response of an isolated bridge depends heavily on the properties of its isolation system,
which in turn derive from the corresponding mechanical properties of the isolators. Part 2 of
Eurocode 8 recognises the following categories of design properties:

Clause 7.5.2.4 [2]

Nominal values of design properties, to be dened by the designer usually in the form of a
range of values characterising an industrial product, such as an isolator. These properties
should be validated in general via special prototype tests. Part 2 of Eurocode 8 gives in
its informative Annex K a description of such tests and the relevant requirements,
applicable if the provisions of EN 15129 (CEN, 2009) are not sufcient and there is no
relevant European Technical Approval (ETA). Prototype tests are not required from
normal low-damping elastomeric bearings (LDEBs) per EN 1337-3 (CEN, 2005b) and
ordinary at sliding bearings per EN 1337-2 (CEN, 2000), whose contributions to an
increase in the damping of the isolation system are ignored.
The variation of these properties due to external factors, such as temperature,
contamination, aging (including corrosion) or wear (expressed as cumulative travel),
should be taken into account in the design, according to normative Annex J of Part 2 of
Eurocode 8. The inuence of these factors is either determined through special tests or, for
common isolator types, estimated on the basis of modication factors (l factors) given in
informative Annex JJ of Part 2 of Eurocode 8. The following two groups of design
properties should be determined and used in separate analysis for each one:
lower-bound design properties (LBDP), usually giving the maximum displacements
upper-bound design properties (UBDP), usually resulting in the maximum forces.
Part 2 of Eurocode 8 gives explicitly, usually in a simplied way, the UBDP values for
normal LDEBs and at sliding bearings.

Annex J,
Annex JJ [2]

7.2.2.5 Restoring capability of isolation system


Large residual displacements of an isolation system after an earthquake should be avoided,
because an offset from the accumulation of such displacements will reduce the displacement
capacity of the system. So, an isolation system should have sufcient automatic restoring
capability (see Section 7.6).
7.2.2.6 Need for sufcient stiffness for non-seismic actions and service conditions
The increased exibility of the horizontal support of the deck due to the isolation system should
not impair the functions of the bridge under service, nor lead to violation of any horizontal
displacement limits set by other Eurocodes for service or ULSs.

Clause 7.7.1 [2]

Clause 7.7.2 [2]

173

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

7.2.3
Conceptual design remarks
The efciency of a seismic isolation solution is measured by the seismic force reduction it entails
in the substructure of the bridge, namely the ratio VI,i/Vf,i , where Vf,i is the seismic shear force of
element i of the xed system and VI,i is the corresponding force of the isolated system, while
keeping the displacements within manageable limits.
As far as the local soil conditions are concerned, the period shift effect referred to in Section 7.2.1
is most efcient if the corner period TC between the constant pseudo-acceleration and the
constant pseudo-velocity ranges of the spectrum is short. Seismic isolation of a bridge
becomes less efcient as the ground type changes from A to B, D, C or E (see Table 3.3 in
Chapter 3). In addition, a longer corner period TD between the constant pseudo-velocity and
the constant displacement ranges of the spectrum adversely inuences the efciency of the isolation. Recall that Part 1 of Eurocode 8 recommends TD 2 s for all ground types; but see,
however, Section 7.3 below.
The effect of the period shift depends also on the value of the fundamental period of the bridge
considered xed, Tf . Bridges with Tf longer than TC and closer to TD do not need seismic isolation, since the forces induced by the seismic action are already low. Examples are bridges on
tall piers, cable-stayed bridges with the deck suspended from the pylon head, etc. Such bridges
may need supplemental damping to reduce seismic displacements and sacricial horizontal tie
systems to control displacements under non-seismic horizontal actions, mainly wind.
A very important favourable side-effect of seismic isolation in bridges is that it very efciently
makes uniform the horizontal seismic forces among piers and abutments of quite different stiffness. Similarly, it reduces the restraint of the deck due to its imposed deformations. This may be
quite important for long continuous post-tensioned concrete decks, which are sensitive to alternating thermal actions. Similar advantages are offered against imposed horizontal displacements
of the foundation, as long as these are small enough to be neglected as far as the displacement
capacity of the isolators is concerned.

7.3.
Clauses 3.2.2.2,
3.2.5(8) [1]
Clauses 3.2.2.3(1),
7.4 [2]

Design seismic action

According to Part 2 of Eurocode 8, the design seismic action of isolated bridges should not be
taken as less than specied for non-isolated bridges. Consequently, whatever has been said in
Chapter 3 regarding elastic spectra, the time history representation of the seismic action, nearsource effects and spatial distribution applies to isolated bridges as well.
Seismically isolated bridges have by design a long fundamental period, and are therefore sensitive
to ground motions rich in low frequencies. Specically, elastic spectra with long values of TC and
TD are very demanding for bridges with seismic isolation. In this respect, a note in Part 2 of
Eurocode 8 allows a value of TD longer than the recommended value of 2 s to be adopted
(intended for earthquakes of magnitude up to 6.5).
Spectral ordinates may increase in the long-period range of the spectrum, and TD may shift up
owing to near-source effects in the vicinity of the seismotectonic fault rupture. The rules in
Part 2 of Eurocode 8 highlighted in Section 3.1.2.6 of this Guide cover the special needs of
bridges with seismic isolation in this respect.

7.4.
Clause 7.5.2.3.2 [2]

Behaviour families of the most common isolators

7.4.1
Bilinear hysteretic behaviour
7.4.1.1 General features
To this family belong many common types of isolators. The most common of them are dealt with
in Sections 7.4.1.2 to 7.4.1.5. Isolators of this family react with a force that depends on the
relative displacement between the isolator interface to the deck and its interface to the supporting
element (pier or abutment). The shape of the cyclic forcedisplacement diagram is very close to
bilinear: unloading branches are parallel to the initial elastic branch at a stiffness (slope) Ke; the
post-yield branches have much lower stiffness (slope) Kp . Figure 7.2 shows a cycle from an
extreme point (Fmax , dbd) to its opposite (Fmax , dbd) and back to (Fmax , dbd). The ratio
Keff Fmax/dbd

174

(D7.1)

Chapter 7. Bridges with seismic isolation

Figure 7.2. Forcedisplacement loop of a bilinear hysteretic system


F
Fmax
Fy
F0

Kp
Ke
dy

Keff

dd

ED
dr = F0 /Kp

termed the effective stiffness, is the stiffness of an elastic system reacting with the same force
Fmax to the peak displacement dbd . To provide the period
Pshift mentioned in Section 7.2, the stiffness of the isolation system, which is equal to the sum
Keff of all isolators, needs be substantially lower than the stiffness of the xed superstructure. The area enclosed by the hysteresis
loop, Ed , is the dissipated energy by the isolator at the peak displacement dbd .
The basic parameters of the bilinear loop are:
g
g

dy and Fy: displacement and force at yield, respectively


F0: force at zero displacement (often termed characteristic strength in the USA),
F0 Fy Kpdy

(D7.2a)

Fmax: force at maximum displacement dbd .

As Fmax and dbd both depend also on the seismic demand, the behaviour of a bilinear system is
intrinsically determined by three parameters: F0 , the elastic stiffness Ke Fy/dy and the postelastic stiffness Kp (see Figure 7.2). Equation (D7.2a) can be used to express Fy and dy as
functions of these parameters:
Fy F0/(1  Kp/Ke)

(D7.2b)

dy (F0/Ke)/(1  Kp/Ke)

(D7.2c)

Other useful relations are


Kp (Fmax  Fy)/(dbd  dy)

(D7.3)

Ed 4F0(dbd  dy) 4(Fydbd  Fmaxdy)

(D7.4)

7.4.1.2 Low-damping elastomeric bearings


LDEBs (depicted in Figure 7.3) offer the possibility of lengthening substantially the fundamental Clause 7.5.2.3.3(2) [2]
period of the system, alongside a practical decoupling of the horizontal response from the inuence of a large stiffness contrast between piers or between them and the abutments. Because in the
early years of application of the pre-standard (ENV) version of Eurocode 8 to bridges, elastomeric bearings were the only devices of European production, widely used, relatively inexpensive
and covered by a European standard that could be used as isolator bearings, Part 2 of Eurocode 8
allows the use of normal LDEBs manufactured per EN 1337-3 (CEN, 2005b) directly as isolators.
Moreover, such bearings have very narrow hysteresis loops and do not lend themselves to
increasing the damping of the bridge above the default value of 5%. Therefore, the usual
elastic response spectrum analysis may be applied with the low shear stiffness of the bearings.
Note, however, that when an elastomer with very low values of the shear modulus is used
(G , 0.6 MPa), scragging may have to be accounted for in determining the design value of G
(see Section 7.4.1.3).
The conceptual rules and verications of these bearings are dealt with in detail in Chapter 6 of
this Guide and their modelling and analysis in Chapter 5. In addition, all the basic principles
175

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.3. Elastomeric bearings: (a) low-damping elastomeric bearing (LDEB); (b) high-damping
elastomeric bearing (HDEB)

and general rules for seismic isolation relating to the present chapter are consistently applicable
to them as well. The following points are noted specically for this type of isolator:
g

The inability of these bearings to offer additional damping increases the system
displacements to the point that they may prove prohibitive for the shear strains of the
bearings and/or the response of the bridge (see Section 6.10.4.4), unless the bearings are
combined with viscous dampers (see Section 7.4.2). This drawback, alongside the fact that
normal LDEBs per EN 1337 that are included in manufacturers lists of standard bearings
are in essence not specically intended for seismic use, makes them inadequate for use in
cases of truly high seismicity or adverse soil conditions.
Although isolation with these LDEBs is feasible only in not very demanding seismic
conditions and cannot be considered as an optimal solution, it is certainly an inexpensive
and simple one. In this respect, damage to some of these bearings in a strong earthquake
may even be considered as acceptable, as long as it is feasible to replace them.

7.4.1.3 High-damping elastomeric bearings


Clause 7.5.2.3.3(3) [2] The damping capability of elastomeric bearings can be increased by replacing the pure elasto-

meric material by special mixtures of an elastomer with special aggregates. The width of the
hysteresis loops increases substantially, and a damping ratio appreciably higher than 5% can
be reached. An isolation system consisting exclusively of such high-damping elastomeric
bearings (HDEBs) may be analysed with the multimodal response spectrum method, if the effective damping ratio of the isolation modes, jeff , is determined from the dissipated energy per cycle,
Ed , as measured by the appropriate tests (see EN 15129 (CEN, 2009)).
Note that, after even a single cycle of shear deformation to the peak shear strain, this isolator type
presents substantial softening (reduction of shear modulus, G) in subsequent cycles (Figure 7.4).
In fact, the virgin value of G is recovered within a few months. This effect, usually termed
scragging, may lead to erroneous determination of the design value of G from tests on
scragged bearings. To avoid this, Part 2 of Eurocode 8 requires that the representative value
of G is determined from tests of unscragged specimens as the average in the rst three cycles
at peak shear strain.
7.4.1.4 Leadrubber bearings
Clause 7.5.2.3.3(9) [2] The most efcient way to achieve substantial energy dissipation capacity of the elastomeric

bearing is by replacing its core by lead (Figure 7.5), which yields early, offering the desired
energy dissipation. The shear stiffness of the lead core and the rubber annulus are

176

KL GLAL/h

(D7.5a)

KR GRAR/hR

(D7.5b)

Chapter 7. Bridges with seismic isolation

Figure 7.4. Forcedisplacement loops of (virgin) HDEBs


50
40
30
Shear force: kN

20
10
0
300

200

100

10

100

200

300

20
30
40
50
Shear strain: %

where GL , AL and h are the shear modulus, the plan area and the thickness of the lead core,
respectively, and GR, AR and hR are those of the rubber annulus, respectively. If FLy is the
yield force of the lead core, the elastic stiffness of the leadrubber bearing (LRB) is
Ke KL KR

(D7.6)

and its yield force is


Fy FLy(1 KR/KL)

(D7.7)

A wide variety of basic parameters of a bilinear system can be achieved by simply varying the
geometry of the lead core and the rubber annulus of the LRB. Many issues concerning the
detailed design of such bearings are given in Constantinou et al. (2011). The values to be
assumed for the material parameters should be in accordance to tests per EN 15129 or to an
applicable European Technical Approval (ETA). UBDP and LBDP bounds should be considered in the design according to Annex J in Part 2 of Eurocode 8. Circular bearings have the
same properties in any horizontal direction, and as such are preferred to square bearings.
7.4.1.5 Steel elastoplastic energy dissipators
These devices are not bearings, as they are not intended to bear gravity loads. Their objective is to
dissipate energy via plastic deformation of ductile steel elements yielding under the horizontal
seismic forces. Such a device is depicted in Figure 7.6(a), while Figure 7.6(b) shows typical
forcedisplacement loops. The restoring capability of systems with such devices may need to
be increased by combining them with other components (e.g. normal LDEBs); alternatively,
the displacement capacity of the system may need to be increased.
7.4.2
Velocity-dependent devices
A velocity-dependent device is connected to the deck and a supporting element (pier or
abutment) at two points, dening its direction of excitation and reaction. It reacts to the
relative motion of the connection points with a force F, which is co-linear to the relative
velocity of these points, v, but in the opposite direction, and has the magnitude
F Cvab

Clause 7.5.2.3.4 [2]

(D7.8)

Exponent ab assumes values in the range 1.000.01, and C is a constant of the device measured by
testing, in units kN/(m/s)ab (Figures 7.7 and 7.8).
Whenever the displacement d becomes a maximum (or absolutely minimum) in a cycle of motion,
and starts decreasing (or increasing), the velocity v passes through zero, as the direction of
motion is reversed. Consequently, the force F also becomes zero. Therefore, for these points
177

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.5. (a) Leadrubber bearings and (b) their forcedisplacement loops at different temperatures

of all cycles, including the maximum, Keff from Eq. (D7.1) is zero. In other words, velocitydependent devices do not contribute to the effective stiffness of the isolation system.
The forcedisplacement relation of velocity-dependent devices can be determined, if the displacementtime relation in a motion cycle is assumed. It is convenient to assume that the displacement
is a sinusoidal function of time t within a cycle of motion with period Teff or angular velocity
v 2p/Teff. Then (see Figure 7.7),
db dbd sin(vt)

(D7.9)

where dbd is the maximum displacement. The velocity v is the time derivative of d:
v vdbd cos(vt)
178

(D7.10)

Chapter 7. Bridges with seismic isolation

Figure 7.6. (a) Steel elastoplastic energy dissipators and (b) typical forcedisplacement loops

and the force from Eq. (D7.8) becomes


F Cvab Fmax[cos(vt)]ab

(D7.11a)

with
Fmax C(dbdv)ab

(D7.11b)

The dissipated energy per cycle at maximum displacement dbd is


ED l(ab)C(dbdv)abdbd

(D7.12)

where the coefcient l(ab) is determined, using the Gamma function G, as

lab 22ab

G2 1 0:5ab
G2 ab

D7:13

The values of l(ab) are given in Table 7.1 for a wide range of ab values.
Figure 7.7. Forcedisplacement loop of velocity dependent devices
F
b = 1

Fmax

b < 1

ED

dd

179

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 7.1. Values of l(a)

a
0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.50 2.00
l(a) 3.988 3.882 3.774 3.675 3.582 3.496 3.416 3.341 3.270 3.204 p
2.876 2.667

Note that the maximum damping force Fmax , as well as the dissipated energy per cycle, Ed ,
depend on v (i.e. on the period of the motion, Teff). The dependence on v is linear and strong
if ab 1.0 (linear damping) or
1.0 and weaker for small values of ab (e.g. 0.01, see
Figure 7.8(b)). Note also that, for any value of ab , the damper force is zero at the point of
maximum displacement, and becomes maximum at zero displacement.
7.4.3
Frictional devices
7.4.3.1 Flat sliding bearings
Friction is a feature common to all sliding bearings. The Coulomb friction law gives the simple
relation between friction force, Ffr , caused by the vertical load, NSd , sliding on a horizontal at
sliding surface (Figure 7.9) with velocity d_b (where d_b is the time derivative of the relative displacement db , i.e. the velocity):
Ffr mfrNSd sign(d_b)

(D7.14)

where mfr is the friction coefcient and sign(d_b) is the sign of the velocity vector, independent of
the magnitude of velocity, depending only on the direction of the motion. However, the value
of mfr depends on the kind and history of the motion. Thus, one may distinguish between:
g
g
g

the dynamic coefcient of friction, md , applicable for seismic motions (i.e. at relatively high
velocities)
the static coefcient of friction ms , representative of very slow quasi-static motion
breakaway friction, for the estimation of the friction force necessary to start sliding after a
longer period of immobility and of load dwell.

Figure 7.8. Velocity-dependent device: (a) a hydraulic viscous damper and (b) its forcevelocity relation
(manufacturer: Maurer-Soehne)
Damper MHD of 3000 kN

3000

Silicon oil

Force: kN

2500
F = C V

2000
1500
1000
500

Piston

Steel cylinder
(a)

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2
Velocity: m/s
(b)

Figure 7.9. (a) A at sliding bearing and (b) its force-displacement loop
F

NSd

F0 = Fmax = dNSd

Fmax
ED
(a)

180

dbd
(b)

Chapter 7. Bridges with seismic isolation

During seismic motion cycles the dynamic value of the friction coefcient mfr md is applicable
and depends on several factors, such as the composition of the sliding surfaces, the use or not of
lubrication, the bearing pressure on the sliding surface, the velocity of sliding, etc.
The friction force remains constant (i.e. Ffr Fmax), therefore Eq. (D7.14) is written as (see also
Figure 7.9)
Fmax mdNSd sign(d_b)

(D7.15)

The dissipated energy per cycle of motion at maximum displacement dbd is


ED 4mdNSddbd

(D7.16)

The above relations apply to an isolator once relative sliding at the sliding surface has started;
that is, when its friction capacity has been reached. This is practically always the case in the
seismic design situation, and can be simply veried by comparing the peak displacement of the
deck with that of the top of the supporting element. For other horizontal actions (wind or
braking load) or imposed deck deformations, certain sliding isolators, especially those supported
on exible piers, may remain inactive if the corresponding connection force demand (with the
connection assumed to be xed) does not exceed the isolator friction capacity.
To use at sliding bearings as energy dissipators in the bridge seismic isolation system, they should
present a reliable lower bound of energy dissipated per cycle, Ed , and, hence, a reliable lower
bound of the friction coefcient md . Normal at sliding bearings conforming to EN 1337-2,
with lubricated PTFE sliding surfaces, may be considered to have a controlled upper bound of
md , but no reliable lower bound under conditions of seismic motion. Therefore, according to
Part 2 of Eurocode 8, these bearings may be used as isolators, but not as energy dissipators.
Clearly, at sliding bearings have no self-restoring capability. So, they may be used as components of an isolation system only in combination with other devices providing the required
restoring capability (e.g. elastomeric bearings).
7.4.3.2 Bearings with one spherical sliding surface
Devices with one spherical sliding surface consist of an articulated slider coated with a special
PTFE material having controlled low friction (Figure 7.10). Sliding occurs on a concave stainless
steel surface with a radius of curvature of the order of 2 m. The coefcient of friction at the sliding
interface is very low, of the order of 0.050.10; it can be reduced even more through lubrication.
The combination of low friction and a restoring force due to the concave surface gives an
approximately bilinear hysteretic behaviour of the bearing, with inherent re-centring capability
(see Figure 7.10). The behaviour consists of the combined effect of:
g

A hysteretic frictional component, which provides the force at zero displacement:


F0 mdNsd

where Nsd is the normal force through the device.


A linear-elastic component that provides a restoring force corresponding to post-yield
stiffness:
Kp

(D7.17)

Nsd
Rb

D7:18

where Rb is the radius of the spherical sliding surface.


The dissipated energy per cycle at the cyclic displacement dbd is as in Eq. (D7.16)
(i.e. ED 4mdNsddbd).

The maximum force, Fmax , and the effective stiffness, Keff , at displacement dbd are
Fmax

 
Nsd
d md Nsd sign d_bd
Rb bd

D7:19
181

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.10. (a) An isolator with one spherical sliding surface and (b) its forcedisplacement loop
dbd
Nsd

Rb

Fmax

Pivot point
(a)
Area enclosed in
loop = dissipated
energy per cycle Ed

Force F

Idealised monotonic response


for static analysis

Fmax
Kp = Nsd /Rb

F0 = dNsd

dbd
Displacement d

Ed

Force-displacement loop
for seismic analysis

(b)

Keff

Nsd md Nsd

Rb
dbd

D7:20

Note that Rb is in fact the radius of the spherical surface on which the pivot point is moving (see
Figure 7.10). If this point lies below the sliding surface at a distance h and that surface is concave
with radius Rs , then Rb Rs h; if it is above (at distance h), then Rb Rs  h.
The elastic stiffness of the equivalent bilinear system Ke is theoretically innite. Numerical
instabilities may be avoided by using Ke F0/dy , with a small value of dy 0.1 mm.
The following special features of these isolators are worth noting:
g

The reaction, Fmax,i , of each isolator i to mass proportional forces acting on the deck is a
horizontal force parallel to the direction of motion and proportional to the corresponding
vertical reaction, Nsd,i . As the forces Nsd,i are in equilibrium with the total weight of the
deck (including overturning action effects in the direction of the motion), the resultant of
the horizontal reactions Fmax,i of all isolators passes through the horizontal projection of
the centre of mass of the deck. Consequently, the seismic motion causes no rotation about
a vertical axis, as long as the deck deformation and the exibility of the piers remain small,
as is the usual case.
Since the resultant horizontal reaction force of these devices is proportional to the mass of
the deck, as is also its inertia force, the dynamic equations of motion become independent
of mass, as long as the mass of the piers may be neglected. Such a system indeed behaves
as a pendulum.

7.4.3.3 Bearings with multiple spherical sliding surfaces


These isolators consist of two external (main) concave plates, within which slide:
g

182

either a spherical bearing with convex external plate surfaces (double spherical slider, see
Figure 7.11 for a schematic) or

Chapter 7. Bridges with seismic isolation

Figure 7.11. Isolator with a double spherical sliding surface


R2, 2

h2

d2
d1

h1

R1, 1
Typically:
R1 = R2 = R; h1 = h2 = h; d1 = d2 = d; 1 = 2 = d

two intermediate convexconcave sliding plates with a central solid convex slider (triple
spherical isolator, shown schematically in Figure 7.12).

The behaviour of the typical double spherical slider has no signicant differences from the simple
one, except that
Rb 2(R  h)

(D7.21)

The displacement capacity, dbd,max , increases to


dlim 2d(1  h/R)

(D7.22)

The behaviour of the triple spherical isolator is more complex. The following concise description
corresponds to the triple friction pendulum bearing, manufactured by EPS. Detailed design procedures for such bearings are given in Constantinou et al. (2011). The cross-section of the generic
type is depicted in Figure 7.12, alongside the usual assumptions for typical bearings. Owing to the
distances hi between the sliding surfaces and the corresponding pivot points, the following
effective value of curvature radii should be used:
Reff,i Ri  hi

(for i 14)

(D7.23a)

di di(Reff,i/Ri)

(for i 14)

(D7.23b)

The following parameters are determined:


d y 2(m1  m2)Reff,2

(D7.24)

m m1  (m1  m2)(Reff,2/Reff,1)

(D7.25)

dlim

d y

2d 1

(D7.26)

Figure 7.12. Triple friction pendulum bearing: geometric and frictional properties
R4, 4

R3, 3
d3

d4

h3

h4

d1

h2

h1

d2
R1, 1

R2, 2

Rigid slider
Slide plates

Typically:
R1 = R4, R2 = R3; h1 = h4, h2 = h3; d1 = d4, d2 = d3; 1 = 4 = d, 2 = 3 < d

183

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.13. Design forcedisplacement loop of a typical triple pendulum bearing

Force F
22Nsd

1Nsd
Nsd

Nsd/2Reff,1
Nsd/2Reff,2

2Nsd
d*y = 2( 1 2)Reff,2

Displacement d
2dy

dbd

Assuming that the maximum displacement demand, dbd , remains less than dlim (dbd  dlim), the
forcedisplacement loops have the trilinear form depicted in Figure 7.13. The energy dissipated is
Ed 4[mdbd  4d y (m  m2)]Nsd

(D7.27)

As a conservative approximation, one may consider the behaviour as bilinear with


dy d y

(D7.28a)

F0 mNsd

(D7.28b)

Kp Nsd/2Reff,1

(D7.28c)

Ke

m1Nsd/d y

(D7.28d)

The behaviour regimes of the device are illustrated in Figure 7.14:


g

Regimes I and II for d y  d  d y (see Figure 7.13) correspond to a force below the
minimum friction m2 or to motion within part of regions d2 and d3 (see Figure 7.12); the
response lies in them during minor seismic events.
Regimes III and IV for d y  d  dlim (see Figure 7.13) correspond to motion within the
main sliding regions d1 and d4 (see Figure 7.12); the response to the design seismic events

Figure 7.14. Triple friction pendulum bearing: behaviour regimes


Total displacement

Regime V
Regime IV

Horizontal force

Regime III
Regime II
Regime I
F

dlim

184

Chapter 7. Bridges with seismic isolation

lies in these regimes. (The second slightly steeper post-yield line reects the general case
when R4 , R1 .)
Regime V reects an additional displacement margin beyond the design displacement
capacity dlim of the device; it corresponds to a motion within the available remaining part
in regions d2 and d3 .

7.5.

Analysis methods

7.5.1
Analysis methods and their elds of application
Part 2 of Eurocode 8 provides for the following analysis methods of bridges with seismic
isolation:
g
g
g

Clause 7.5.3 [2]

nonlinear time-history analysis (NLTH)


the fundamental mode method (FMM)
multimode spectrum analysis (MMS).

The rst one is the most general analysis method, and may be applied to all cases. The conditions under which the two spectral methods may be applied in a stand-alone way are given
in Table 7.2.
7.5.2
Nonlinear time-history analysis
As the superstructure and the substructure remain essentially elastic, their stiffness is in general
established considering them as uncracked; the nonlinearity of the model should reect the
nonlinear properties of the isolation system, including:
g
g

Clauses 7.5.6(1),
7.5.7(1) [2]

the interaction of the simultaneous response in at least the two horizontal directions
the effect of overturning forces.

Regarding specically the contribution of the vertical seismic component, alternatively to using a
simultaneous vertical acceleration time history, Part 2 of Eurocode 8 allows its estimation via the
linear response spectrum method and combination of the action effects with those of the horizontal components via Eq. (D5.2c).
Note that the energy dissipation offered by the components of the isolation system, either as
hysteretic damping (through bilinear hysteretic or friction devices) or as viscous (via hydraulic
dampers), is directly accounted for in a nonlinear analysis. It is therefore important to ensure
that the input data given for the damping matrix of the system, intended to represent the
standard structural damping of the substructure and superstructure elements, are not interpreted
by the computational algorithm as applicable to the isolated modes as well (i.e. for the modes
with the longest period). It should also be taken into account that the main part of the seismic
deformation energy in the structure refers to these modes. On the other hand, as always in
nonlinear analysis, it is desirable to use increased damping to lter out possibly inaccurate
results due to very short period modes (i.e. those with a period less than the time-step of the
direct time integration which is usually of the order of 0.01 s). Usually, Raleigh damping is
applied to meet such limitations, with the damping matrix C derived as a linear combination
of the stiffness matrix K and the mass matrix M:
C aK bM

(D7.29)

Then, for any natural period T,

j ap/T bT/4p

(D7.30)

Table 7.2. Conditions for the application of spectral methods


Spectrum method

Fundamental mode method


Multimodal analysis

Conditions for applicability


Distance from active fault

Ground type

Effective damping, jeff

10 km
No limit

A, B or C
A, B or C

0.30
0.30

185

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

The above mentioned requirements can be satised by the selection

b0

a jT/p 0.10  0.05/p 0.00159

(D7.31)

which gives (see also Figure 8.52 in Chapter 8)


g
g

Clause 7.5.4 [2]

for T  1.5 s: j  0.0032


for T  0.05 s: j  0.10.

7.5.3
Fundamental mode method of analysis
In the fundamental mode method (Figure 7.15) the entire structure is taken as a single-degree-offreedom (SDoF) system according to the rigid deck model of Sections 5.6.3 and 5.6.4. In its basic
form it considers only the stiffness and dissipation energy properties of the isolation system in
one horizontal direction at a time. Despite its simplied form, the method can capture the
most signicant aspects of the structural response. Certain corrections and adaptations are
discussed in Section 7.5.5.
The method estimates the maximum seismic displacement dcd of the system iteratively, using two
basic approximation tools:
g

The effective stiffness of the system, Keff . The nonlinear forcedisplacement relation of the
isolation (the system is shown schematically in Figure 7.15(a)) is obtained by summing up
the force contributions of all its components, i, corresponding to an assumed displacement
value dcd,a . The effective stiffness is the secant stiffness at the point of maximum
displacement, dcd,a:
Keff Fmax/dcd,a

Fmax,i/dcd,a

(D7.32)

Keff,i

Any velocity-dependent devices do not contribute to Keff (see Section 7.4.2).


The effective damping of the system, jeff: This is the equivalent viscous damping
corresponding to the sum of dissipated energies, ED,i , of all components of the isolation
system at the cycle of peak displacement dcd dcd,a . It is calculated as

jeff

P

ED;i
1

2
2p Keff dcd

D7:33

The values of ED,i are known isolator properties in terms of the isolator displacement.
The deck seismic displacement is taken as that of a linear SDoF system having a mass equal to
that of the deck, Md, stiffness Keff and an equivalent viscous damping ratio jeff . Its period is
Teff

s
Md
2p
Keff

D7:34

Figure 7.15. Fundamental mode method: (a) forcedisplacement; (b) displacement spectra
= 0.05

Keff

= eff

Fmax
dcd.r

dcd.a
(a)

186

TC

Teff TD
(b)

Chapter 7. Bridges with seismic isolation

The peak displacement demand, dcd,r , can be derived as shown schematically in Figure 7.15(b),
from the displacement response spectrum that results from multiplication of the elastic acceleration spectrum for j 0.05 by the damping modication factor, heff , corresponding to the
estimated value of jeff:
s
0:10
heff
D7:35
0:05 jeff
After a few iterations, the assumed value dcd,a converges to the seismic demand dcd,r .
This simple approach gives good estimates of the maximum seismic forces, only if they occur
simultaneously with the estimated maximum displacement. This holds in general if the
reacting force of all isolators of the system is either constant (friction) or depends only on the
displacement (as in elastic or hysteretic isolators). Then,
Fmax ,i Keff,idcd

(D7.36)

When the isolation system also contains devices with a reacting force proportional to velocity
(hydraulic viscous dampers), their maximum force does not take place at the instant of peak
displacement. In such cases, a more complex procedure is required to estimate the peak forces
(see Section 7.5.5.5).
7.5.4
Multimode spectrum analysis
This is an approximate method, combining via the usual combination rules of Section 5.5.4:
1
2

Clause 7.5.6 [2]

the fundamental mode method described above for the isolation modes,
the multimodal response spectrum analysis with the default damping ratio of j 0.05 for
the higher modes.

As this is an approximation, it is preferable to apply it separately in the two horizontal directions


(longitudinal and transverse) and to combine the results via Eqs (D5.2). The same applies for the
combination with the computed effects of the vertical component.
Step 2 (a multimodal analysis for the upper modes) is applied as follows in each horizontal
direction:
g
g
g

All isolators i are modelled with their effective stiffness, Keff,i , as derived from the
fundamental mode method in the direction considered.
The substructure and superstructure elements are modelled with uncracked stiffness and a
sufcient number of intermediate nodes to capture the effects of important higher modes.
The modal damping ratio is taken as equal to the default value j 0.05 for all modes with
a period T , 0.8Teff , where Teff is the effective period of the isolated mode from the
fundamental mode method; the effective damping of that mode, jeff , is used for all modes
with periods longer than 0.8Teff .

Application of this method is meaningful only to capture possibly signicant contributions of


higher modes. The limitation pointed out in Section 7.5.3 regarding the inability to estimate
well the maximum forces in combinations of viscous dampers and hysteretic isolators still holds.
7.5.5

Effects of isolation properties on the approximation of the fundamental


mode method results
7.5.5.1 Introduction
Part 2 of Eurocode 8 and its counterparts (e.g. AASHTO, 2010) use the results of the fundamental mode method as lower limits for the results of the multimode spectral method or even
the nonlinear time-history approach, as a sort of self-checking by the designer. It is therefore
important to control the approximation it offers and to improve it wherever possible.
The following deals briey with the impact of Keff , jeff and bidirectional excitation on the
approximation of the fundamental mode method and offers suggestions for its improvement
187

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.16. Composite pier stiffness (Part 2 of Eurocode 8)


did
dbi,d

Deck
Fi
Isolator i

Hi

Pier i

Fi /Kti

Fi /Ksi

Fi /Kbi

Fi Hi2/Kri

where possible. It also describes methodologies for estimating the maximum forces in isolation
systems using viscous dampers

Clause 7.5.4(3) [2]

7.5.5.2 Effective stiffness Keff


The basic form of the fundamental mode method accounts only for the stiffness of the isolators,
neglecting the exibility of the piers. Although this assumption is acceptable in most usual cases,
it is not for high and exible piers.
The exibility of an abutment and its foundation is usually negligible. The composite stiffness of
pier i may be obtained from Eq. (D2.10), rewritten here as (Figure 7.16)
1
Keff;i

1
1
1
H2

i
Kbi Kti Ksi Kfi

D7:37

where Kbi , Kti and Ksi are the translational stiffness of the bearing, the foundation and the pier
shaft, respectively, Kfi is the foundation rotational stiffness and Hi the height of the pier shaft.
The last three summands are lumped into the pier exibility 1/KPi . Then,
Keff,i KbiKPi/(Kbi KPi)

(D7.38)

If Kbi/KPi is small, Eq. (D7.38) may be written as


Keff,i Kbi/(1 Kbi/KPi)  Kbi(1  Kbi/KPi)

(D7.39)

If the isolators over all piers and abutments have the same stiffness, Kbi Kb , then,
Keff

Keff,i

Kb  K2b

(1/KPi)

(D7.40)

This correction of Eq. (D7.32) separately in each horizontal direction lends itself to hand
calculation.
7.5.5.3 Damping ratio jeff
Only few of the forcedisplacement loops from an analysis of either an isolated or a non-isolated
ductile bridge reach a displacement near the peak displacement demand among these loops, in
the response to an accelerogram used for the design. Most other loops have substantially
smaller peak displacements. However, the value of jeff corresponding to the loop with the
maximum displacement is used in the fundamental mode method for the estimation of the
peak displacement demand. It is therefore informative to study the variation of jeff as a
function of the displacement in a given loop.
Figure 7.17 shows the variation of jeff in a bilinear hysteretic system as a function of the
displacement normalised by the yield displacement (the displacement ductility ratio m d/dy)
and the hardening ratio, l Kp/Ke . Figures 7.18 and 7.19 depict the evolution of the ratio
188

Chapter 7. Bridges with seismic isolation

Figure 7.17. Variation of the effective damping ratio jeff of a bilinear hysteretic system as a function of
the ductility ratio m d/dy for various values of l Kp/Ke
0.60

0.001
0.01

0.50

0.05

0.30

eff

0.40

0.10
0.15

0.20
0.10
0.00
0

7 8 9 10 11 12 13 14 15
= d/dy

jeff/jeff,c of effective damping with the displacement amplitude d of the loop normalised to the
peak displacement during the response, dc . Figure 7.18 concerns bilinear hysteretic systems
with the maximum ductility ratio mc dc/dy as the parameter, whereas Figure 7.19 refers to a
system of elastic isolators (e.g. LDEB) combined with viscous dampers, and is parametrised
by the damper exponent a.
It is clear from Figures 7.17 and 7.18 that at ductility levels around 10 and higher, jeff either
remains approximately constant with displacement or even reduces slightly at larger displacement amplitudes. The reduction with the displacement amplitude is more pronounced in
Figure 7.19 for viscous dampers with small a values. For a 1.0 (linear damping), jeff
remains constant over the whole range of d/dc up to 1.0. For this reason, the use of effective
damping, jeff , gives satisfactory results in cases of seismic isolation, because the ductility ratio
demands reached are high. By contrast, Figures 7.17 and 7.18 show that for bilinear hysteretic
systems reaching ductility ratios mc less than 57 (as typical of reinforced concrete or steel
non-isolated ductile structures), jeff is smaller at low displacement amplitudes. More rened
hysteretic models for reinforced concrete structures exhibit similar trends (Fardis and Panagiotakos, 1996). The large scatter of experimental results shown also in Fardis and Panagiotakos
(1996) is an additional reason why effective damping, jeff , cannot be reliably applied for
estimating nonlinear displacement demands in ductile reinforced concrete structures.
7.5.5.4 Bidirectional excitation
The fundamental mode method refers to SDoF systems and motions in a certain direction.
However, real isolation systems have, in general, two horizontal degrees of freedom, with
Figure 7.18. Evolution of the ratio jeff/jeff,c of effective damping to its value at the peak displacement dc
with the ratio d/dc of the peak displacement of a loop d to dc, in a bilinear hysteretic system, as a
function of the nal ductility ratio mc dc/dy, for a hardening ratio l Kp/Ke equal to 0.05
1.6
1.4

c = 12.5

c = 10

c = 7.5

1.2
eff /eff,c

1.0

c = 15

0.8
0.6

c = 5

= 0.05

0.4
c = 3

0.2
0.0
0.0

0.1

0.2

0.3

0.4

0.5
d/dc

0.6

0.7

0.8

0.9

1.0

189

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 7.19. Evolution of the ratio jeff/jeff,c of effective damping to its value at the peak displacement dc
with the ratio d/dc of the peak displacement in a loop d to dc, in an elastic system with only viscous
damping as a function of the damper exponent a
3.0

= 0.40

= 0.00
= 0.05
= 0.10
= 0.20

2.5

eff /eff,c

2.0

= 0.60

1.5

= 0.80

1.0

= 1.00

0.5
0.0
0.0

0.1

0.2

0.3

0.4

0.5
d/dc

0.6

0.7

0.8

0.9

1.0

usually (but not always) the same or similar properties in these two directions. Part 2 of
Eurocode 8, as well as AASHTO (2010) and more recently Constantinou et al. (2011), do not
explicitly provide for the estimation of the effect of the transverse seismic action in the context
of the fundamental mode method. In fact, Eqs (D5.2) normally underestimate this effect.
Clauses 3.2.3(3),
3.2.3(6), 3.2.3(7) [2]

The key to the modication of the fundamental mode method to make its results compatible with
those of nonlinear time-history analysis lies in the rules governing the compatibility of the pairs
of horizontal acceleration histories with the elastic response spectrum, highlighted in Section
3.1.4 of this Guide. These rules lead to the following options for achieving compatibility
between the fundamental mode method and nonlinear time-history analysis:
The fundamental mode method is conveniently carried out using an elastic spectrum
multiplied by a factor of 1.251.30. This amplication slightly exceeds the requirement of
Eurocode 8.
2 The results of the fundamental mode method carried out for each horizontal component
strictly according to Part 2 of Eurocode 8 are modied by multiplying the displacements
by a factor between 1.15 and 1.25. This amounts to assuming a concurrent transverse
displacement between 57% and 75% of the estimated peak displacement.
1

Clause 7.6.3(9) [2]

7.5.5.5 Estimation of maximum forces in systems with viscous dampers


For isolation systems consisting of elastic isolators (e.g. elastomeric bearings) and exclusively
viscous dampers, Part 2 of Eurocode 8 gives a methodology for the estimation of maximum
forces occurring simultaneously on elastic isolators and dampers. However, it does not give
any guidance for systems of hysteretic isolators and viscous dampers. A general theoretical
solution for free vibrations of such a system is given in Ribeiro et al. (2007). In this section,
the more practical methodology in Constantinou et al. (2011) is highlighted.
Both the hysteretic isolators with dissipated energy per cycle ED,h from Eq. (D7.3) and the
viscous dampers with dissipated energy ED,v from Eqs (D7.12) and (D7.13) contribute to the
effective damping as

jeff (

ED,h

P
ED,v)/(2p Keffd 2cd)

(D7.41)

Designating by jv the part of jeff contributed by viscous damping,


P
P
jv ( ED,v)/(2p Keffd 2cd)
190

(D7.42)

Chapter 7. Bridges with seismic isolation

Table 7.3. Correction factor rv


Effective
period: s

Effective damping: %
10

20

30

40

50

60

70

80

90

100

0.3
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0

0.72
0.75
0.82
0.95
1.08
1.05
1.00
1.09
0.95

0.70
0.73
0.83
0.98
1.12
1.11
1.08
1.15
1.05

0.69
0.73
0.86
1.00
1.16
1.17
1.17
1.22
1.15

0.67
0.70
0.86
1.04
1.19
1.24
1.25
1.30
1.24

0.63
0.69
0.88
1.05
1.23
1.30
1.33
1.37
1.38

0.60
0.67
0.89
1.09
1.27
1.36
1.42
1.45
1.49

0.58
0.65
0.90
1.12
1.30
1.42
1.50
1.52
1.60

0.58
0.64
0.92
1.14
1.34
1.48
1.58
1.60
1.70

0.54
0.62
0.93
1.17
1.38
1.54
1.67
1.67
1.81

0.49
0.61
0.95
1.20
1.41
1.59
1.75
1.75
1.81

the maximum velocity vm in sinusoidal motion can then be estimated from the pseudo-velocity
vdcd as
vm rvvdcd

(D7.43)

The correction factors rv are given in Table 7.3 (see Constantinou et al., 2011). The maximum
forces Fm on the system can be estimated as


2pjv
ab
ab
Fm Keff dcd cos d
rv sin d
D7:44
 Keff dcd
l
where


2p aj v
d
l

1=2ab

D7:45

and the other variables are dened in Eqs (D7.13) and (D7.42).

7.6.

Lateral restoring capability

Part 2 of Eurocode 8 requires the isolation system present in both horizontal directions sufcient
lateral restoring capability to automatically prevent accumulation of large residual displacements. Note that the absence of residual displacements in the results of nonlinear time-history
analyses using few (usually 310) horizontal pairs of spectrum-compatible ground acceleration
histories is by no means conclusive proof of sufcient restoring capability. The criteria in Part
2 of Eurocode 8 have a solid basis: several hundred thousand nonlinear analyses (Katsaras
et al., 2008) of bilinear hysteretic systems subjected to an exhaustive range of strictly natural
far- and near-source records.

Clause 7.7.1 [2]

REFERENCES

AASHTO (2010) Guide Specications for Seismic Isolation Design. American Association of State
Highway and Transportation Ofcials, Washington, DC.
CEN (Comite Europe de Normalisation) (2000) EN 1337-2:2000: Structural bearings Part 2:
Sliding elements. CEN, Brussels.
CEN (2004) EN 1998-1:2004: Eurocode 8 Design of structures for earthquake resistance Part 1:
General rules, seismic actions and rules for buildings. CEN, Brussels.
CEN (2005a) EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance Part 2:
Bridges. CEN, Brussels.
CEN (2005b) EN 1337-3:2005: Structural bearings Part 3: Elastomeric bearings. CEN, Brussels.
CEN (2009) EN 15129:2009: Antiseismic devices. CEN, Brussels.
Constantinou MC, Kalpakidis I, Filiatrault A and Ecker Lay RA (2011) LRFD-based Analysis and
Design Procedures for Bridge Bearings and Seismic Isolators. Department of Civil, Structural and
Environmental Engineering, State University of New York, Buffalo, NY. Technical Report
MCEER-11-004:2011.
191

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Fardis MN and Panagiotakos TB (1996) Hysteretic damping of reinforced concrete elements.


11th World Conference on Earthquake Engineering, Acapulco, Paper 464.
Katsaras CP, Panagiotakos TB and Kolias B (2008) Restoring capability of bilinear hysteretic
seismic isolation systems. Earthquake Engineering and Structural Dynamics 37(4): 557575.
Ribeiro AMR, Maia NMM, Fontul M and Silva JMM (2007) Complete Response of a SDoF
System with a Mixed Damping Model. Departamento de Engenharia Mecanica, Instituto
Superior Tecnico, Lisbon.

192

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance


ISBN 978-0-7277-5735-7
ICE Publishing: All rights reserved
http://dx.doi.org/10.1680/dber.57357.193

Chapter 8

Seismic design examples


8.1.

Introduction

This chapter presents examples of seismic design of bridges according to Part 2 of Eurocode 8.
The generic example is a bridge having a continuous deck with one central span and two
shorter side spans. The chapter comprises the following examples:
g

g
g

Section 8.2 example of ductile piers: a bridge with a concrete deck monolithic with piers
designed for ductile behaviour. This earthquake-resisting system is usually cost-effective
for bridges of medium-short spans and medium total length.
Section 8.3 example of limited ductile piers: a longer bridge on tall piers designed for
limited ductile behaviour.
Section 8.4 example of seismic isolation: a longer bridge on short piers with seismic
isolation.

All three examples have developed from the contribution of the senior author of this Designers
Guide (Bouassida et al., 2012).

8.2.

Example of a bridge with ductile piers

8.2.1
Bridge layout design concept
The bridge is a three-span overpass, with spans 23.50 35.50 23.5 m and a total length of
82.5 m. The deck is a post-tensioned cast-in-situ concrete voided slab. The piers consist of
single cylindrical columns with diameter D 1.2 m, monolithic with the deck. The pier heights
are 8 m for M1 and 8.5 m for M2. The bridge is simply supported on the abutments through a
pair of bearings allowing free sliding and rotation in every horizontal direction. The piers and
the abutments are supported on piles. The concrete grade is C30/37. The conguration of the
bridge and cross-sections of the deck and the piers are shown in Figures 8.18.3.
The selection of single cylindrical-column piers allows arranging the supports at right angles to
the longitudinal axis, despite the slightly skew crossing in plan. For the given geometry of the
bridge, the monolithic connection between the piers and the deck minimises the use of expensive
bearings or isolators (and their maintenance), without subjecting the bridge elements to excessive
Figure 8.1. Longitudinal section of the bridge with ductile piers
A1

M1

M2

A2

193

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.2. Plan view of the bridge with ductile piers


A1

M1

M2

A2

restraint by imposed deck deformations. Some comments on the cost-effectiveness of the seismic
resistant system are given, as conclusions, in Section 8.2.12.
8.2.2
The earthquake-resisting system
8.2.2.1 Structural system and ductility
The main elements resisting seismic forces are the piers. A ductile seismic behaviour is selected for
these elements. The value of the behaviour factor q, as given in Section 5.4, Table 5.1, depends on
the shear span ratio Ls/h of the piers:
g

For the longitudinal direction (taken as the horizontal direction X), assuming the piers
to be fully xed to the foundation and to the deck and for the shortest pier M1:
Ls 8.0/2 4.0 m and Ls/h 4.0/1.2 3.33 . 3.0; therefore:, qX 3.50.
For the transverse direction (taken as the horizontal direction Y), assuming the piers to be
fully xed to the foundation and free to move and rotate at the connection to the deck and
for the shortest pier M1: Ls 8.0 m and Ls/h 8.0/1.2 6.67 . 3.0, allowing qY 3.50.

8.2.2.2 Stiffness of elements


8.2.2.2.1 Piers
The value of the effective stiffness of the piers for the seismic analysis is estimated initially and
checked after dimensioning the vertical reinforcement of the piers. For both piers the stiffness
is assumed to be 40% of the uncracked exural stiffness of the gross section.
8.2.2.2.2 Deck
The full uncracked exural stiffness of the gross section of the prestressed concrete deck is taken.
Considering the voided section as closed, the torsional stiffness is 50% of the uncracked gross
section stiffness.
Figure 8.3. Cross-sections of (a) the piers and (b) the deck

(a)

194

(b)

Chapter 8. Seismic design examples

Figure 8.4. Design response spectrum (normalised to the design peak ground acceleration, ag)
0.900
0.821
0.800
0.767
0.700

Regions corresponding to constant:


Velocity

Acceleration

Displacement

Sd(T )/ag

0.600
0.500
0.400
0.300
0.200

0.100
TC

TB
0.000
0.00

0.50

TD
1.00

1.50
Period, T: s

2.00

2.50

3.00

8.2.3
Design seismic action
A response spectrum of type 1 applies for the design seismic action. The ground type is C, with
the recommended values of the soil factor S 1.15 and the corner periods TB 0.2 s and
TC 0.6 s in Table 3.3, whereas the corner period is taken as longer: TD 2.5 s. The bridge is
located at a seismic zone with a reference peak ground acceleration agR 0.16g. The importance
factor is gI 1.0, and the design peak ground acceleration in the horizontal directions is
ag gIagR 1.0  0.16g 0.16g
The lower bound factor for design spectral accelerations is b 0.20. For the behaviour factors
qX 3.5 in the longitudinal direction and qY 3.5 in the transverse direction, the design
response spectrum (normalised to the design peak ground acceleration, ag) from Eqs (D5.3) in
Section 5.3 is shown in Figure 8.4.
8.2.4
Quasi-permanent actions for the seismic design situation
The loads applied to the bridge deck (Figure 8.5) in the seismic design situation are as follows.
8.2.4.1 Self-weight (G)
The area of the voided section is 6.89 m2, the area of the solid section is 9.97 m2, the total length
of the voided section is 73.5 m and the total length of the solid section is 9.0 m:
qG (6.89 m2  73.5 m 9.97 m2  9.0 m)  25 kN/m3 14 903 kN
8.2.4.2 Additional dead load (G2)
The area of the sidewalks is 0.50 m2/m, the weight of the safety barriers is 0.70 kN/m and the
width and thickness of the pavement are 7.5 m and 0.10 m, respectively:
qG2 2  25 kN/m3  0.50 m2 (sidewalks) 2  0.70 kN/m (safety barriers) 7.5 m
 23 kN/m3  0.10 m (road pavement) 43.65 kN/m
Figure 8.5. Dead, additional dead and quasi-permanent uniform trafc load application

195

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

8.2.4.3 Quasi-permanent value of trafc load (LE)


The quasi-permanent value of the trafc load is taken as 20% of the uniformly distributed trafc
load (UDL) of load model 1 (LM1), qUDL 45.2 kN/m:
qQP 0.20qL 0.2  45.2 kN/m 9.04 kN/m
8.2.4.4 Thermal actions (T)
The thermal actions consist of:
g

a uniform temperature difference DTN,ext 52.58C of the maximum uniform bridge


temperature component, Te,max , from the initial temperature T0 108C at the time the
deck is erected;
a uniform temperature difference DTN,con 458C of the minimum uniform bridge
temperature component, Te,min , from the initial temperature, T0, at the time the deck is
erected.

No vertical temperature difference component, DTM, between the top and bottom surfaces of the
deck is considered.
8.2.4.5 Creep and shrinkage (CS)
A total strain of 32.0  10  5 is considered. It is of relevance only for the bearing displacements.
The total load applied on the bridge deck in the seismic design situation is then
WE 14 903 kN (43.65 9.04) kN/m  82.5 m 19 250 kN
8.2.5
Fundamental mode analysis in the longitudinal direction
The fundamental mode period is estimated based on a simplied single-degree-of-freedom
(SDoF) cantilever model of the bridge. The mode corresponds to the oscillation of the bridge
along its longitudinal axis, assuming both ends of the piers as xed.
For cylindrical columns of diameter 1.2 m, the uncracked moment of inertia is Iun p1.24/64
0.1018 m4. The assumed effective moment of inertia of piers is Ieff/Iun 0.40 (to be checked
later). Assuming both ends of the piers xed, for concrete grade C30/37 with Ecm 33 GPa,
the stiffness of each pier in the longitudinal direction is
K1 12EIeff/H3 12  33 000 MPa  (0.40  0.1018 m4)/(8.0 m)3 31.5 MN/m
K2 12EIeff/H3 12  33 000 MPa  (0.40  0.1018 m4)/(8.5 m)3 26.3 MN/m
The total longitudinal stiffness is:
K 31.5 26.3 57.8 MN/m
The total seismic weight is: WE 19 250 kN, so the fundamental period is
r
r
M
19 250=9:81
T 2p
2p
1:16 s
K
57 800
The spectral acceleration in the longitudinal direction is
Se agS(2.5/q)(TC/T) 0.16g  1.15  (2.5/3.5)  (0.60/1.16) 0.068g
The total seismic shear force in the piers is
VE SeWE/g 0.068g  19 250 kN/g 1309 kN
The shear force is distributed to piers M1 and M2 in proportion to their stiffness:
V1 (31.5/57.8)  1309 kN 713 kN
V2 1309  713 596 kN
196

Chapter 8. Seismic design examples

Table 8.1. Normal modes


Mode No.

1
2
3
4
5
6
8
15
16
17
18
26
27
28
30

Period: s

1.77
1.43
1.20
0.32
0.32
0.19
0.15
0.054
0.053
0.052
0.051
0.030
0.029
0.028
0.027

Modal mass contribution: %


X direction

Y direction

Z direction

0
0
99.2
0
0
0
0
0
0
0
0
0.2
0
0
0.1

3.4
94.8
0
0
0.3
0.7
0
0
0
0
0
0
0.1
0
0

0
0
0
8.9
0
0
63.1
10.8
0.2
1.8
0.4
0.2
0
5.4
0.1

The longitudinal seismic moments My (assuming full xity of pier columns at top and bottom)
are
My1  V1H1/2 713 kN  8.0 m/2 2852 kN m
My2  V2H2/2 596 kN  8.5 m/2 2533 kN m
8.2.6
Modal response spectrum analysis
The nite element discretisation of the deck is very ne, the same as used for the analysis for
gravity and trafc loads. The piers are also discretised vertically in a fairly large number of
elements.
The characteristics of the 15 most important modes of the structure out of a total of 50 computed
in the analysis are shown in Table 8.1. Some of these modes, as well as those not listed in
Table 8.1, have negligible contribution to the total response. The shapes of the most important
among the rst eight modes are depicted in Figure 8.6.
Response spectrum analysis considering the rst 50 modes was carried out. The sum of the modal
masses considered amounts to 99.6%, 99.7% and 92% in the X, Y and Z directions, respectively.
The combination of modal responses was carried out using the complete quadratic combination
(CQC) rule, Eqs (D5.20) and (D5.22b) in Section 5.5.4. The results of the fundamental mode
analysis in the longitudinal direction and of the modal response spectrum analysis are presented
and compared in Table 8.2.
8.2.7
Design action effects and verications
8.2.7.1 Design action effects for exure and axial force verication of plastic hinges
The combination of the components of the seismic action is carried out according to Eqs (D5.2)
in Section 5.2. Table 8.3 gives the design action effects (bending moment and axial force) at
the bottom section of pier M1, together with the required reinforcement, for each design
combination.
The pier is of circular section with diameter D 1.2 m, of concrete C30/37 with Class C steel of
S500 grade. The nominal cover is c 50 mm, and the estimated distance of the bar centre from
the surface is 82 mm. The required reinforcement at the bottom section of pier M1, which is
critical, is 19 870 mm2. The reinforcement selected is 25 132 (20 100 mm2), as shown in
Figure 8.7. Figure 8.8 displays the MomentAxial force interaction diagram for the bottom
section of Pier M1 for all design combinations.
197

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.6. Modes 1, 2, 3, 4 and 8

Table 8.4 shows the design action effects of bending moment and axial force at the bottom
section of pier M2 together with the required reinforcement, for each design combination.
The required reinforcement at the bottom section of Pier M2, which is critical, is 16 800 mm2.
The reinforcement selected is 21 132 (16 880 mm2) as shown in Figure 8.9. Figure 8.10 shows
the momentaxial force interaction diagram for the bottom section of pier M2 for all design
combinations.

Table 8.2. Comparison of analyses in the longitudinal direction


Pier

Fundamental mode analysis:


1.16 s

Multimode response spectrum analysis:


1.20 s (third mode)

Seismic shear, Vz

M1
M2

713 kN
596 kN

662 kN
556 kN

Seismic moment, My

M1
M2

2852 kN m
2533 kN m

At top: 2605 kN m. At bottom: 2672 kN m


At top: 2327 kN m. At bottom: 2381 kN m

Fundamental period, T1

198

Chapter 8. Seismic design examples

Table 8.3. Design action effects and required reinforcement at the bottom section of pier M1
Combination

N: kN

My: kN m

Mz: kN m

As: mm2

max My Mz
min My Mz
max Mz My
min Mz My

7159
7500
7238
7082

4576
3720
713
456

1270
1296
4355
4355

19 870
13 490
17 240
17 000

8.2.7.2 Conrmation of stiffness of ductile piers


The effective stiffness of the piers for the seismic design situation is estimated according to Eqs
(D5.33)(D5.35) in Section 5.8.1.
8.2.7.2.1 Pier M1
For one layer of 25 132 (20 100 mm2) and an axial force of N 7200 kN, using the design
values of material properties, the design value of yield moment is My 4407 kN m; the steel
yield strain is
1sy fyd/Es 500/(1.15  200 000) 0.00217
Figure 8.7. Pier M1 cross-section with reinforcement

Figure 8.8. Momentaxial force interaction diagram of the bottom section of pier M1
30 000
Section with 2532
25 000
20 000
Design
combinations

15 000
N: kN

10 000

6000

5000
4000

2000

2000

4000

6000

5000
10 000
15 000
M: kN m

199

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 8.4. Design action effects and required reinforcement at the bottom section of pier M2
Combination

N: kN

My: kN m

Mz: kN m

As: mm2

max My Mz
min My Mz
max Mz My
min Mz My

7528
7145
7317
7320

3370
4227
465
674

1072
1042
3324
3324

10 320
16 800
8980
9250

while the corresponding strain at the extreme concrete bres is 1cy 0.00272. The design value of
the moment resistance (again for design values of material properties) is MRd 4779 kN m. The
yield curvature is

fy (0.00217 0.00272)/(1.2  0.082) 4.37  10  3 m  1


while the approximation of Eq. (D5.35b) for circular sections yields

fy 2.41sy/d 2.4  0.00217/(1.2  0.082) 4.66  10  3 m  1


g

Equation (D5.33a) in Section 5.8.1 yields


Ic p  1.24/64 0.1018 m4
My/(Ecfy) 4407/(33 000  4.37  10  3) 0.0306 m4
Ieff 0.08 Ic Icr 0.0387 m4
(EI )eff/(EI )c 0.38

Equation (D5.33b) in Section 5.8.1 gives


(EI )eff 1.2MRd/fy 1.2  4779/4.37  10  3 1 312 000 kN m2
Ieff 1 312 000/33 000 0.0398 m4
(EI )eff/(EI )c 0.39

The assumed value (EI )eff/(EI )c 0.40 was a good estimate for the analysis.
8.2.7.2.2 Pier M2
For pier M2 the nal reinforcement is one layer of 21 132 (16 880 mm2). The axial force is
N 7200 kN. For the design values of material properties the yield moment is
My 4048 kN m, the steel yield strain is 1sy 0.00217 and the corresponding strain at the
Figure 8.9. Pier M2 cross-section with reinforcement

200

Chapter 8. Seismic design examples

Figure 8.10. Momentaxial force interaction diagram of the bottom section of pier M2
30 000
Section with 2132
25 000
20 000

N: kN

15 000

Design
combinations

10 000
5000
0

6000

4000

2000

2000

4000

6000

5000
10 000
M: kN m

extreme concrete bres 1cy 0.00273. The design value of the moment resistance is
MRd 4366 kN m.
From Eq. (D5.33a) in Section 5.8.1: (EI )eff/(EI )c 0.35.
From Eq. (D5.33b) in Section 5.8.1: (EI )eff/(EI )c 0.36.

g
g

The assumed value (EI )eff/(EI )c 0.40 was a fairly good estimate for the analysis.
8.2.7.3 Dimensioning of the piers in shear
8.2.7.3.1 Overstrength moments
The overstrength moment is calculated as Mo goMRd , where go is the overstrength factor and
MRd is the design value of moment resistance (see Section 6.4.1, Eqs (D6.6)). The normalised
axial force is

hk NEd/Ac fck 7600/(1.13  30) 0.22


Since hk 0.22 . 0.1, the minimum value go 1.35 is multiplied by
1 2(hk  0.1)2 1 2(0.22  0.1)2 1.029
So,

go 1.35  1.029 1.39


and the overstrength moments for the piers are:
g
g

Mo1 1.39  4779 6643 kN m


Mo2 1.39  4366 6069 kN m

8.2.7.3.2 Pier capacity design shears


In the longitudinal direction the shear forces from the analysis in piers M1 and M2 are
V1 713 kN and V2 596 kN, while the capacity shears can be calculated directly from the
overstrength moments:
g
g

VC1 2Mo1/H1 2  6643/8.0 1661 kN


VC2 2Mo2/H2 2  6069/8.5 1428 kN

In the transverse direction, the shear in each pier is calculated applying the simplications of
Annex G of Part 2 of Eurocode 8 (see Section 6.4.2):
VCi (Mo/MEi)VEi
201

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.11. Capacity design action effects in the transverse direction

VC2 = 1251 kN
Mo2 = 6069 kN m
Mo1 = 6643 kN m

VC1 = 1476 kN

The seismic moments and shear forces are:


g
g

ME1 3061 kN m and VE1 680.3 kN


ME2 2184 kN m and VE2 450.2 kN

So, the capacity design shear forces are


g
g

Vc1 (6643/3061)  680.3 1476 kN


Vc2 (6069/2184)  450.2 1251 kN

Figure 8.11 shows the capacity design action effects in the transverse direction at the base of the
piers.
8.2.7.3.3 Dimensioning in shear
For a circular section
d r rs 0.60 0.52 1.12 m
while the effective depth is
de r 2rs/p 0.60 2  0.52/p 0.93 m
The shear strength of the section is calculated as
VRd,s (Asw/s)(0.9de) fywd cot u/gBd or as VRd,s (p/4)(Asw/s)(0.9d ) fywd cot u/gBd
where Asw is the total cross-sectional area of the shear hoops, s is their spacing, fywd is their design
yield strength, u is the angle between the concrete compression strut and the pier axis, with
cot u 1 as per Part 2 of Eurocode 2 and gBd is the safety factor in note 5 of Table 6.1, with
the value gBd 1.25  (qVEd/VC,o  1)  1.
For pier M1 the design shear force is VC1 1661 kN, and that from the analysis is V1 713 kN.
Then,
1.25  (qVEd/VC,o  1) 1.25  (3.5  713/1661  1) 0.75
so gBd 1 and the required shear reinforcement is
Asw/s 1.0  1661/(0.84  0.5/1.15  1.0) 4550 mm2/m
or
Asw/s 1.0  1661/(0.7854  0.9  1.12  0.5/1.15  1.0) 4827 mm2/m
For pier M2 the design shear force is VC1 1428 kN and that from the analysis is V1 596 kN.
Then,
1.25  (qVEd/VC,o  1) 1.25  (3.5  596/1428  1) 0.79
202

Chapter 8. Seismic design examples

so gBd 1 and the required shear reinforcement is


Asw/s 1.0  1428/(0.84  0.5/1.15  1.0) 3910 mm2/m
or
Asw/s 1.0  1428/(0.7854  0.9  1.12  0.5/1.15  1.0) 4150 mm2/m
8.2.8
Ductility requirements for the piers
8.2.8.1 Connement reinforcement
The normalised axial force is

hk NEd/Ac fck 7600/(1.13  30 000) 0.22 . 0.08


so connement of the compression zone is required (see note 15 of Table 6.1).
The longitudinal reinforcement ratio is:
g

for pier M1:

rL 20 100/1 130 000 0.0178


for pier M2:

rL 16 880/1 130 000 0.0149


The distance from the surface to the spiral centreline is c 58 mm (Dsp 1.084 m), and the core
concrete area is Acc 0.923 m2.
For circular spirals and ductile behaviour, the required mechanical reinforcement ratio, vw,req , of
connement reinforcement is (see Table 6.1) vw,req 0.52(Ac/Acc)hk 0.18( fyd/fcd)(rL  0.01),
and vw,min 0.18. For
g

pier M1:

vw,req 0.52  (1.13/0.923)  0.22 0.18  (500/1.15)/(0.85  30/1.5)


 (0.0178  0.01) 0.176
g

pier M2:

vw,req 0.52  (1.13/0.923)  0.22 0.18  (500/1.15)/(0.85  30/1.5)


 (0.0149  0.01) 0.162
For the worst case of pier M1:

vwd,c max(vw,req; vw,min) 0.18


and the required volumetric ratio of conning reinforcement is

rw vwd,c( fcd/fyd) 0.18  (0.85  30/1.5)/(500/1.15) 0.0070


while the required area of conning reinforcement (one leg) is
Asp/sL rwDsp/4 0.007  1.084/4 0.0019 m2/m 1900 mm2/m
The maximum allowed hoop spacing is (see Table 6.1):
max sL 1084 mm/5 217 mm
203

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 8.5. Comparison of transverse reinforcement requirements of piers


Requirement

Connement

Buckling of bars

Shear design

At/sL: mm2/m
max sL: mm

2  1900 3800
217

164

M1: 4825, M2: 4150

8.2.8.2 Prevention of buckling of longitudinal bars


For the transverse reinforcement required to prevent buckling of the longitudinal bars, the
maximum hoop spacing sL should not exceed ddbL , where

d 2.5( ftk/fyk) 2.25  5


(see Table 6.1). For S500 steel, ftk/fyk
1.15:

d 2.5  1.15 2.25 5.125


Therefore,
max sL ddL 5.125  32 mm 164 mm
8.2.8.3 Transverse reinforcement of piers comparison of requirements
The pier transverse reinforcement requirements for each design check are compared in Table 8.5.
The transverse reinforcement is governed by shear design. The reinforcement selected for both
piers is one spiral of 116/85 (4730 mm2/m).
8.2.9
Capacity design verications of the deck
8.2.9.1 Estimation of capacity design effects an alternative procedure
The general procedure in Part 2 of Eurocode 8 for calculating the capacity design effects consists
of adding to the action effects of the quasi-permanent loads G, the effects of the loading
DAC Mo  G, both acting in the deck-cum-piers frame system of the bridge (see Section
6.4.2 of this Guide).
An alternative procedure for the longitudinal (X) direction is to work on a continuous beam
system of the deck, simply supported on the piers and the abutments. On this system the
effects of the quasi-permanent loads G and those of the overstrength moments Mo are
added. Figure 8.12 demonstrates the equivalence of the two procedures.
The effects of the quasi-permanent loads G are displayed in Figure 8.13. Figure 8.14 shows the
effects of the overstrength loading Mo for seismic action in the X direction. For the effects due
Figure 8.12. Equivalence of general and simpler capacity design procedures for the deck
Permanent load G
G

MG1

MG2

MG1

MG2

MG1

MG2

MG1

MG2

MG1

MG2

MG1

MG2

Ac: Over strength G


MO1 MG1

MO2 MG2

MO1 MG1

MO2 MG2

MO1 MG1

MO2 MG2
General procedure

204

MO1

MO2
MO1

MO1
;

MO2
MO2

alternative procedure

Chapter 8. Seismic design examples

Figure 8.13. Quasi-permanent loads (G loading) and resulting moment and shear force diagrams
G

M: kN m

2195
2285

2430

4052

2340

V: kN
3207

(+)

()

3211

4047

to seismic action in X direction, the signs of the effects are reversed. Figure 8.15 displays the
result of the superposition of these two loadings to get the capacity effects for seismic action
in the X direction.
8.2.9.2 Flexural verication of the deck
The deck section at each side of the connection of the deck with a pier is checked against the
capacity design effects, taking into account the existing reinforcement and tendons (shown in
Figure 8.16). Table 8.6 lists the moment and axial force combinations for which the deck
sections are checked. Figure 8.17 compares the momentaxial force ultimate limit state (ULS)
interaction diagram with the capacity design effects.
8.2.9.3 Other deck verications
The ULS verication of the deck in shear is, in general, not critical. So, it is not presented
here.
The verication of the connection between the pier and the deck as a beam/column joint
per Section 6.4.4 is far from critical, and therefore not presented here. It is usually critical
for the shear reinforcement of joints with slender pier columns monolithically connected to the
deck.
Figure 8.14. Overstrength for seismic action in the X direction (Mo loading): resulting moment and
shear force diagrams
Mo1 = 6643 kN m

Mo2 = 6069 kN m

M: kN m
4296

3692

3104

4491

6069

6643
6069

6643

V: kN
()
187.4

256.7

263.8

(+) 1661

161.9

(+) 1428

205

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.15. Capacity effects for seismic action in the X direction (G Mo): resulting moment and
shear force diagrams
M: kN m
6491

6122

764

2206

6069

6643
6069

6643

V : kN
3795

3045
(+)
()
3398

(+)

4311

1661

(+) 1428

Figure 8.16. Deck section, reinforcement and tendons


Top layer: 46 20 + 33 16 (210.8 cm2)

4 groups of 3 tendons
of type DYWIDAG 6815
(area of 2250 mm2 each)

Bottom layer: 2 33 16 (182.9 cm2)

8.2.10 Design action effects for the foundation design


Figure 8.18(a) displays the capacity design effects acting on the foundation of pier M1 for the
longitudinal direction, for seismic actions in the negative direction (X). Figure 8.18(b) shows
the capacity design effects for the transverse direction. The sign of the effects is reversed for
the opposite direction of the seismic action in the transverse direction.

Table 8.6. Loading combinations for the deck section

206

Connection to

Direction of seismic action

My: kN m

N: kN

Pier
Pier
Pier
Pier
Pier
Pier
Pier
Pier

X
X
X
X
X
X
X
X

6122
2206
6491
764
1262
6776
2101
5444

29 900
28 300
29 500
28 100
28 100
29 500
28 300
29 900

M1
M1
M2
M2
M1
M1
M2
M2

left side
right side
left side
right side
left side
right side
left side
right side

Chapter 8. Seismic design examples

Figure 8.17. Momentaxial force interaction diagram of the deck section


200 000
Design combinations
150 000

N: kN

100 000

50 000

0
60 000

40 000

20 000

20 000

40 000

50 000
M: kN m

Figure 8.18. Capacity design effects on the foundation of pier M1: (a) seismic action in the longitudinal
(X) direction; (b) seismic action in the transverse direction
N = 7371 kN

N = 7371 kN

Vz = 1661 kN

Vz = 730 kN

Vy = 1476 kN

My = 6643 kN m
X

My = 124 kN m
Mz = 6643 kN m

Y
(a)

(b)

8.2.11 Bearings and roadway joints


8.2.11.1 Bearings
The design displacement of the bearing, dEd , is given by Eq. (D6.36).
The displacements in the longitudinal direction are shown in Figure 8.19(a), and those in the
transverse direction in Figure 8.19(b). The maximum displacement at the bearings is 93.9 mm
and 110 mm, respectively.
The bridge is simply supported on the abutments through a pair of bearings allowing free
sliding and rotation for both horizontal axes. Plan and side views of a bearing are depicted in
Figure 8.20.
Figure 8.19. Bridge displacements (mm): (a) in the longitudinal direction; (b) in the transverse direction

207

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.20. Plan and side views of sliding bearings


Plan

40
140

460
360

95.5

Longitudinal
axis

500

600

1/2 view; 1/2 section

140

440
540
40

Uplifting of the bearings is checked. The minimum vertical reaction forces in the bearings are
presented in Figure 8.21, with a minimum resultant value of 17.8 kN (compressive, so there is
no uplifting). The maximum vertical reactions are shown in Figure 8.22, with a maximum
value of 2447 kN.
8.2.11.2 Overlapping length
The minimum overlapping (seating) length at moveable joints is given by Eq. (D6.34), where:
g
g
g

the support length is in the present case equal to lm 0.5 m . 0.4 m


the effective seismic displacement of the support from the analysis for the seismic design
situation is in the present case equal to des 0.101 m
for the effective ground displacement deg (2dg/Lg)Leff:
the design ground displacement is in the present case
dg 0.025agSTCTD 0.025  0.16  9.81  1.15  0.6  2.5 0.068 m
the distance parameter for ground type C is Lg 400 m
the effective length of the deck is in the present case Leff 82.5/2 41.25 m.

So, for no proximity to a known seismically active fault:


deg (2dg/Lg)Leff (2  0.068/400)  41.25 0.014 m , 2dg 0.136 m
Figure 8.21. Minimum reaction forces in the bearings in the seismic design situation (kN)

208

Chapter 8. Seismic design examples

Figure 8.22. Maximum reaction forces in the bearings in the seismic design situation (kN)

Therefore,
min lov 0.50 0.014 0.101 0.615 m
The available seating length is 1.25 m . min lov .
8.2.11.3 Roadway joints
The width of a roadway joint between the top deck slab and the top of the backwall of the
abutment should be designed to accommodate the displacement given by Eq. (D6.39). The clearance between the deck structure and the abutment or its backwall should accommodate the larger
displacement given by Eq. (D6.36).
Table 8.7 gives the displacements for the roadway joint and those for the clearance of the structure. Due to the differences between the two clearances, Part 2 of Eurocode 8 requires detailing of
the backwall to cater for predictable (controlled) damage. An example of such a detail is shown in
Figure 8.23, where impact along the roadway joint is foreseen to occur on the approach slab.
Figure 8.24 displays the selected roadway joint type and the displacement capacities for each
direction.
8.2.12 Conclusions regarding the design concept
Optimal cost-effectiveness of a ductile bridge system is achieved when all its ductile elements
(notably the piers) have dimensions that lead to a seismic demand that is critical for the main
reinforcement of all critical sections and exceeds the minimum reinforcement requirements.
This is difcult to achieve when the piers resisting the earthquake have:
g
g

substantial height differences or


sections larger than those required for the purposes of seismic design.

In such cases, it may be more economical to use:


g
g

limited ductile behaviour if the design peak ground acceleration, ag, is low or
exible connection to the deck (seismic isolation).

Table 8.7. Displacement for the roadway joint and clearance at the joint region
Displacement: mm

dG

dT

Longitudinal
Opening
Closure

18.5
0

10.5
8.5

Transverse

dE

dEd,J ( joint)

dEd (structure)

76
76

54.5
34.5

100.5
80.5

+110

+44.0

+110

209

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.23. Clearances and detailing of the roadway joint region


Clearance of
roadway joint
Approach slab
Structure
clearance

Figure 8.24. Selected roadway joint type

Roadway joint type: T120


Capacity in longitudinal direction: 60 mm
Capacity in transverse direction: 50 mm

It is worth noting that Part 2 of Eurocode 8 does not specify a minimum reinforcement requirement (see note 23 in Table 6.1). For the bridge in this example, the owner specied rmin 1%
(a fairly high value), as required by Part 1 of Eurocode 8, but only for the columns of buildings.
The longitudinal reinforcement of the piers is derived from the seismic demands, and was over the
minimum requirement (rL 1.78% for pier M1 and rL 1.49% for pier M2).

8.3.

Example of a bridge with limited ductile piers

8.3.1
Bridge layout design concept
A schematic of the bridge is depicted in Figure 8.25. The deck is straight, continuous, with spans
of 60 80 60 m. It has a composite (steel, concrete) section, consisting of two built-up steel
girders connected at regular intervals via built-up cross bracings and of a concrete slab.
The piers are of reinforced concrete, 40 m tall. Their section is constant throughout the pier
height, hollow circular, with a thickness of 0.4 m and an external diameter of 4 m. The two
girders are supported on the pier through bearings; each pair of bearings is supported on a
pier head that is 4 m wide and 1.5 m deep. Concrete class in the piers is C35/45, and the steel
is of grade S500, Class B.

Figure 8.25. Bridge elevation and arrangement of the bearings


Pinned
connection

Pinned
connection
Sliding longitudinal
pinned transverse
Y

C0R

P1R

P2R

C3R
X

C0L

210

P1L

P2L

C3L

Chapter 8. Seismic design examples

The large exibility of the 40 m tall piers has the following structural consequences:
g
g

The connection of the deck to both pier heads can be pinned (hinged) about the transverse
axis, without causing excessive restraints due to imposed deck deformations.
The large exibility of the seismic force-resisting system gives long fundamental periods in
both horizontal directions and quite low response spectral accelerations. For such low
seismic response levels it is neither expedient nor cost-effective to design the piers for
ductility. Therefore, a limited ductile behaviour is selected with behaviour factor q 1.5
(see Table 5.1).

8.3.2
Design seismic action
A response spectrum of type 1 applies. The ground type is B, with the recommended values of the
soil factor S 1.2 and of the corner periods TB 0.15 s, TC 0.50 s and TD 2 s in Table 3.3.
The bridge is in a seismic zone with reference peak ground acceleration agR 0.3g. The
importance factor is gI 1, and the design peak ground acceleration in the horizontal directions
ag gIagR 1.0  0.3g 0.3g. The lower bound factor for design spectral accelerations is
b 0.2. The behaviour factor is q 1.5. The design response spectrum from Eqs (D5.3) in
Section 5.3 is shown in Figure 8.26.
8.3.3
Seismic analysis
8.3.3.1 Quasi permanent trafc loads
The quasi permanent value c2.1Qk,1 of the UDL system of LM1 is applied in the seismic design
situation. For bridges with severe trafc, the value of c2,1 is 0.2.
The load of the UDL system of LM1, with adjustment factors aq 1.0 for the UDL, are:
g

lane 1:

aqq1,k 3 m  9 kN/m2 27.0 kN/m


g

lane 2:

aqq2,k 3 m  2.5 kN/m2 7.5 kN/m


g

lane 3:

aqq3,k 3 m  2.5 kN/m2 7.5 kN/m

Figure 8.26. Design spectrum for horizontal components with q 1.5


0.7

Ag = 0.3, ground type = B, soil factor = 1.2


Tb = 0.15 s, Tc = 0.5 s, Td = 2.0 s, = 0.2

0.6

Acceleration: g

0.5
0.4
0.3
0.2
0.1
0
0

0.5

1.0

1.5

2.0

2.5

3.0 3.5
Period: s

4.0

4.5

5.0

5.5

6.0

211

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.27. Structural model

residual area:

aqqr,k 2 m  2.5 kN/m2 5.0 kN/m


Total load 47.0 kN/m.
In the seismic design situation the trafc load applied per unit of length of the bridge is

c2,1Qk,1 0.2  47.0 kN/m 9.4 kN/m


8.3.3.2 Structural model
The structural model employs prismatic 3D beam elements for cross-bracings and the two girders
of the deck. Prismatic elements are also used for the piers and in the pier heads (Figure 8.27). The
piers were considered xed at the top of the foundation. The model of each bearing takes into
account the pertinent constraints between the DoFs of the deck and pier head nodes connected
by the bearing.
Figure 8.28. Momentcurvature curve of a pier section with r 1%
75 000
70 000

Bottom section

65 000
60 000

Top section

55 000
Moment: kN m

50 000
45 000
= 1%

40 000
35 000
30 000
25 000
20 000
15 000

N = 15 000 kN
N = 20 000 kN

10 000
5000
0.0
0

212

1 103

2 103

3 103
Curvature

4 103

5 103

6 103

Chapter 8. Seismic design examples

Figure 8.29. Moment(EI)eff /(EI )c ratio curve of a pier section with r 1%


75 000
70 000

N = 15 000 kN
N = 20 000 kN

65 000
60 000
55 000
Bottom section

Moment: kN m

50 000
45 000
40 000

= 1%

35 000
30 000
Top section

25 000
20 000
15 000
10 000
5000
0.0
0.00

0.10

0.20

0.30

0.40

0.50
J/Jgross

0.60

0.70

0.80

0.90

1.00

8.3.3.3 Effective pier stiffness


The effective pier stiffness was initially assumed as 50% of the uncracked gross section stiffness,
(EI)c , giving fundamental periods of 3.88 s in the longitudinal (X) direction and 3.27 s in the
transverse direction (Y). The lower bound of the design spectrum in Eqs (D5.3c) and (D5.3d)
of Section 5.3, equal to Sd bag 0.2  0.3g 0.06g, applies for T  3.3 s. So, the exact
value of (EI )eff is immaterial for the design seismic action effects, provided that
(EI )eff , 0.5(EI)c . However, the analysis was carried out in the end with (EI )eff 0.3(EI)c , to
avoid underestimating the displacements. As shown in Figures 8.288.31, in the range of axial
forces (1520 MN) and bending moments (about 60 MN m) of interest here, this value corresponds better to the nally required reinforcement ratio of r 1.5%. With this (EI )eff value,
the fundamental period in the longitudinal direction is 5.02 s, and in the transverse direction it
is 3.84 s.

Moment: kN m

Figure 8.30. Momentcurvature curve of a pier section with r 1.5%


95 000
90 000
85 000
80 000
75 000
70 000
65 000
60 000
55 000
50 000
45 000
40 000
35 000
30 000
25 000
20 000
15 000
10 000
5000
0.0

Bottom section

Top section

= 1.5%

N = 15 000 kN
N = 20 000 kN
0

1 103

2 103

3 103
Curvature

4 103

5 103

6 103

213

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Moment: kNm

Figure 8.31. Moment(EI)eff /(EI )c ratio curve of a pier section with r 1.5%
95 000
90 000
85 000
80 000
75 000
70 000
65 000
60 000
55 000
50 000
45 000
40 000
35 000
30 000
25 000
20 000
15 000
10 000
5000
0.0
0.00

N = 15 000 kN
N = 20 000 kN

Bottom section
= 1.5%

Top section

0.10

0.20

0.30

0.40

0.50
J/Jgross

0.60

0.70

0.80

0.90

1.00

8.3.3.4 Response spectrum analysis


The characteristics of the 11 most important modes out of the total of 30 computed in the modal
analysis are shown in Table 8.8. Some of them, as well as those not listed in Table 8.8, have
negligible contribution to the total response.
The shapes of four modes are presented in Figures 8.328.35.
A response spectrum analysis considering the rst 30 modes was performed. The sum of modal
masses considered amounts to 97.1% and 97.2% in the X and Y directions, respectively. The
combination of modal responses was carried out using the CQC rule.
Figure 8.36 shows the distribution of peak bending moments along pier P1.

Table 8.8. Normal modes


No.

1
2
3
4
6
9
11
17
18
23
27
28

214

Period: s

5.03
3.84
1.49
0.79
0.66
0.48
0.42
0.26
0.26
0.16
0.15
0.13

Modal mass: %
X direction

Y direction

Z direction

92.5
0.0
0.0
0.0
0.0
0.0
0.0
0.0
4.4
0.0
0.0
0.0

0.0
76.8
0.0
0.0
8.4
2.1
0.0
6.2
0.0
0.0
3.0
0.0

0.0
0.0
0.0
1.2
0.0
0.0
63.2
0.0
0.0
5.0
0.0
8.8

Chapter 8. Seismic design examples

Figure 8.32. First mode longitudinal

Figure 8.33. Second mode transverse

8.3.4
Second-order effects
8.3.4.1 Geometric imperfections of the piers and second-order effects
The inclination of the pier according to clause 5.2 of EN 1992-2:2005 (CEN, 2005a) is
p
u uoah 1/200  2/ l, where l is the length or height of the pier (l 40 m). Therefore,
3
u 1.58  10 rad. According to clause 5.2(7) of EN 1992-1-1:2004 (CEN, 2004a), this
inclination creates an eccentricity, ei uilo/2, where lo is the effective length:
g
g

longitudinal direction, X: (lo 80 m) ex 0.063 m


transverse direction, Y: (lo 40 m) ey 0.032 m.

Figure 8.34. Third mode rotation

215

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.35. Eleventh mode vertical

The eccentricities under permanent loads (G), including the creep effect (for a creep coefcient
w 2.0), are amplied for second-order effects according to Eq. (D6.4) in Section 6.3.1 with
n NB/NEd (NB is the buckling load and NEd the axial force). The results are shown in
Table 8.9.
8.3.4.2 Second-order effects of the seismic action
These effects are estimated using two approaches.
(i) According to clause 5.8 of EN 1992-1-1:2004
The nominal stiffness method (clause 5.8.7 in Part 1-1 of Eurocode 2) is applied using
(EI )eff 0.3(EI), compatible with the seismic design situation. The moment magnication
Figure 8.36. Peak bending moment distribution in the two directions of the section of pier P1

.813
6678304.6

6227
6
829..287
1

26
43.93.3
1313 24

8
86373.0
.193

0
969.494.1
191893

1111334
405.2.831

6
738.4
6564.5
2622

1144223
475.8.087

70.67
1299.59
3273

117733
1232.5.355

30.25
5946.98
393

20206

.88
5847
24068.3
6
4

24241
4165.6
8.6
24

53

216

.13
237 0.89
67
3
5

1612.7
2.1
02

272
87688.2
13.86

Chapter 8. Seismic design examples

Table 8.9. First- and second-order eccentricities due to geometric imperfections of the piers
Direction

ei: m

n NB/NEd

ei,II/ei

ei,II: m

X
Y

0.063
0.032

19.65
78.62

1.161
1.039

0.073
0.033

factor is evaluated at the bottom section as 1 b/[(NB/NEd)  1], where b 1, NEd is the design
value of axial load (19 538 kN) and NB is the buckling load based on the nominal stiffness
NB p2(EI )eff/(b1L0)2, with b1 1. This gives the following moment magnication factors:
g
g

1.154 in the longitudinal direction X


1.034 in the transverse direction Y.

(ii) According to EN 1998-2:2005


The increase in the bending moments at the plastic hinge section (self-weight of the pier included)
is given by the rst term in Eq. (D6.3b) in Section 6.3.1: DM (1 q)dEdNEd/2, where dEd is the
seismic displacement of the pier top and NEd is the axial force from the seismic analysis given in
Table 8.10. This approach from Part 2 of Eurocode 8 (CEN, 2005b) gives approximately the
same moments in the longitudinal direction as that of Eurocode 2, but much higher values in
the transverse direction. It is used in the further design calculations in Table 8.11.
8.3.5
Action effects for the design of piers and abutments
Table 8.10 lists the action effects of the individual actions and of their combinations for the
seismic design situations. The effects are given:
g
g

for piers P1 and P2, at the base section


for abutments C0 and C3 at the mid-distance between the bearings.

The designation of the individual actions is as follows:


g
g
g
g
g
g

G
EX
EY
P-D EC2
P-D EC8
Imperf

Permanent quasi-permanent trafc loads


Earthquake in the X direction (for the design spectrum with q 1.5)
Earthquake in the Y direction (for the design spectrum with q 1.5)
Additional second-order effects according to clause 5.8 of EN 1992-1-1:2004
Additional second-order effects according to EN 1998-2:2005
First- and second-order effects (including creep) of geometric pier imperfections.

8.3.6
Action effects for the design of foundation
Table 8.12 gives the action effects corresponding to the loading combinations of the seismic
design situation for the design of the foundations. The seismic action effects correspond to
q 1.0. The action effects are given:
g
g

for piers P1 and P2, at the base section


for abutments C0 and C3 at the mid-distance between the bearings

with the designation depicted in Figure 8.37. The signs of shear forces and bending moments
given are mutually compatible. However, as these effects (with the exception of the vertical
axial force Fz) are due predominantly to the seismic action, their signs and senses may be
reversed.
Table 8.10. Pier top displacements dEd
Pier top displacement

dx: mm

dy: m

EX 0.3EY
EY 0.3EX

0.373
0.110

0.065
0.197

217

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 8.11. Action effects for the design of piers and abutments: q 1.50 ((EI )eff 0.3(EI)c)
Actions or combination thereof

Fx Vx:
kN

Fy Vy:
kN

Fz N:
kN

5.4
5.4
0.0
0.0

0.2
0.2
0.2
0.2

19539.3
19539.3
3505.2
3505.2

6.8
6.8
0.4
0.4

216.9
216.9
0.0
0.0

0.2
0.2
0.0
0.0

Mx :
kN m

My:
kN m

Mz:
kN m

P1
P2
C0
C3

G
G
G
G

P1
P2
C0
C3

EX 0.3EY
EX 0.3EY
EX 0.3EY
EX 0.3EY

1254.4
1254.4
0.0
0.0

187.4
187.4
322.2
322.2

28.8
28.8
21.5
21.5

7885.5
7885.5
1134.3
1134.3

50803.5
50803.5
0.0
0.0

342.8
342.8
0.0
0.0

P1
P2
C0
C3

EY 0.3EX
EY 0.3EX
EY 0.3EX
EY 0.3EX

376.3
376.3
0.0
0.0

624.6
624.6
1073.9
1073.9

8.6
8.6
6.4
6.4

26285.1
26285.1
3781.1
3781.1

15241.0
15241.0
0.0
0.0

1142.5
1142.5
0.0
0.0

P1
P2
P1
P2

EX 0.3EY P-D
EX 0.3EY P-D
EY 0.3EX P-D
EY 0.3EX P-D

EC2
EC2
EC2
EC2

1254.4
1254.4
376.3
376.3

187.4
187.4
624.6
624.6

28.8
28.8
8.6
8.6

8153.6
8153.6
27178.8
27178.8

58576.4
58576.4
17572.9
17572.9

342.8
342.8
1142.5
1142.5

P1
P2

EX 0.3EY P-D EC8


EX 0.3EY P-D EC8

1254.4
1254.4

187.4
187.4

28.8
28.8

9279.0
9279.0

58298.5
58298.5

342.8
342.8

P1
P2
P1
P2

EY 0.3EX P-D EC8


EY 0.3EX P-D EC8
EX 0.3EY P-D EC8 Imperf
EX 0.3EY P-D EC8 imperf

376.3
376.3
1254.4
1254.4

624.6
624.6
187.4
187.4

8.6
8.6
28.8
28.8

30508.3
30508.3
9826.3
9826.3

17451.4
17451.4
59393.1
59393.1

1142.5
1142.5
342.8
342.8

P1
P2

EY 0.3EX P-D EC8 Imperf


EY 0.3EX P-D EC8 Imperf

376.3
376.3

624.6
624.6

8.6
8.6

31055.6
31055.6

18546.0
18546.0

1142.5
1142.5

P1
P2
C0
C3

G EX 0.3EY P-D EC8 Imperf 1259.8


G EX 0.3EY P-D EC8 Imperf 1249.1
G EX 0.3EY
0.0
G EX 0.3EY
0.0

187.2
187.2
322.3
322.3

19568.0
19568.0
3526.6
3526.6

9833.1
9833.1
1134.0
1134.0

59610.0
59176.2
0.0
0.0

342.5
343.0
0.0
0.0

P1
P2
C0
C3

G EY 0.3EX P-D EC8 Imperf


G EY 0.3EX P-D EC8 Imperf
G EY 0.3EX
G EY 0.3EX

624.4
624.4
1074.1
1074.1

19547.9
19547.9
3511.6
3511.6

31062.4
31062.4
3780.8
3780.8

18762.9
18329.1
0.0
0.0

1142.3
1142.7
0.0
0.0

381.7
371.0
0.0
0.0

Table 8.12. Action effects for the design of the foundation: q 1.0 ((EI )eff 0.3(EI)c)

218

Actions or combination thereof

Fx:
kN

Fy:
kN

P1,
P2

G EX 0.3EY P-D EC8 Imperf


G EY 0.3EX P-D EC8 Imperf

1887.0
569.8

C0,
C3

G EX 0.3EY
G EY 0.3EX

0.0
0.0

Fz :
kN

Mx :
kN m

My:
kN m

Mz:
kN m

280.9
936.7

19582.4
19552.2

13775.8
44204.9

85011.7
26383.4

513.9
1713.5

483.4
1611.1

3537.4
3514.8

1701.1
5671.3

0.0
0.0

0.0
0.0

Chapter 8. Seismic design examples

Figure 8.37. Positive sense and direction of forces and moments for the foundation design
Vertical
direction Z
Mz
Longitudinal
direction X

Transverse
direction Y
My

Fz

Mx
Fx

Fy

8.3.7
Verication of the piers
8.3.7.1 ULS in exure and axial force
Dimensioning of the reinforcement at the base section is as follows: the nominal cover is
c 50 mm and the estimated distance of the bar centre from the surface is 80 mm. For
design action effects, NEd 19 568 kN, My 59 610 kN m and Mx 9833 kN m ! As,req
67 800 mm2. The selected longitudinal reinforcement is:
g
g

at the exterior perimeter: 62 128 (38 100 mm2)


at the interior perimeter: 49 128 (30 100 mm2).

Figure 8.38 depicts the interaction diagram for the design of the base section.

Figure 8.38. Interaction diagram for the design of the pier base section
40 000
= 1.5%
= 1.0%

30 000
20 000
10 000
0
10 000
20 000
Axial force: kN

30 000
40 000
50 000
60 000
70 000
80 000
90 000
100 000
110 000
120 000
130 000
140 000

10 000

20 000

30 000

40 000

50 000

60 000

70 000

80 000

Moment: kN m

219

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

8.3.7.2 ULS in shear


For limited ductile behaviour the pier seismic shear force from the analysis is multiplied by
q 1.5 (see Section 6.3.2 and Table 6.1), giving
p
VEd 1:5 1259:22 187:22 1910 kN
while the resistance, VRd, is divided by gBd1 1.25 (see note 4 in Table 6.1).
The shear resistance at diagonal compression failure of a member with annular section is
VR;max

p
t D  2cnfcd sin 2d
4 w

with 0.4  tan d  1 (228  d  458)


and




fck MPa
35
n 0:6 1 
0:6 1 
0:516
250
250
VR;max

1 p
0:85  35 000
 0:4  4:0  2  0:08  0:516 
sin 2d
1:25 4
1:5

9877 sin 2d
6811 kN > VEd
after division by gBd1 1.25 and for the lower limit of tan d 0.4.
Neglecting the contribution of the axial load to shear resistance as small in this slender and lightly
loaded pier, that of the transverse steel, VRs , is
VRs

p Asw
f D  2c cot d
2 sh ywd

after division by gBd1 1.25, the lower limit value of 0.4 for tan d gives
VRs

1 p Asw 0:5
A
4  0:16  2:5  VEd ! sw  182 mm2 =m
sh
1:25 2 sh 1:15

8.3.8
Ductility requirements
8.3.8.1 Conning reinforcement
The normalised axial force is

hk 19 560/(35 000  4.52) 0.1236 . 0.08


so, connement is required (see note 15 in Table 6.1).
For circular spirals and limited ductile behaviour the required mechanical reinforcement ratio,
vw,req , of connement reinforcement is (see Table 6.1)

vw,req 0.39(Ac/Acc)hk 0.18( fyd/fcd)(rL  0.01)


and

vw,min 0.12
vw,req (4.52/3.39)  0.39  0.1236 0.18(500/1.15)/(0.85  35/1.5)
 (0.015  0.01) 0.085.
220

Chapter 8. Seismic design examples

So, the mechanical reinforcement ratio is

vwd,c max(0.085; 0.12) 0.12


and the required volumetric ratio of conning reinforcement is

rw vwd,c( fcd/fyd) 0.12  (0.85  35/1.5)/(500/1.15) 0.005474


The required area of conning reinforcement (one leg) is
Asp/sL rwAcc/(pDsp) 0.005474  3.39/(p  3.384) 0.001746 m2/m 1746 mm2/m
8.3.8.2 Prevention of buckling of longitudinal bars
The maximum hoop spacing, sL, of transverse reinforcement to prevent buckling of vertical bars
is ddbL , where d 2.5( ftk/fyk) 2.25  5 (see Table 6.1). For Class B steel, ftk/fyk  1.08, and

d 2.5  1.08 2.25 4.95


So d 5, and
sreq
L ddL 5  28 mm 140 mm
As pointed out in Section 4.4.1.5 of this Guide, at the inside face of annular piers, hoops are not
efcient in preventing buckling of vertical bars or in conning the concrete, as they cannot offer
tensile hoop action. If vertical bars yield in compression or the crushing strain of unconned
concrete is exceeded at the inside face under the design seismic action (which is not the case in
this example), cross-ties as for straight boundaries are necessary. 116/110 (1828 mm2/m) is
nally chosen.
8.3.9
Bearings and joints
Table 8.13 gives the seismic deformation and force demands on the bearings. An example of the
design for overlapping length at the movable supports and for roadway joints has been given in
Sections 8.2.11.2 and 8.2.11.3, respectively.

8.4.

Example of seismic isolation

8.4.1
Introduction
This section covers the design of the bridge in the example in Section 8.3 but with a special seismic
isolation system capable of resisting high seismic loads. The seismic isolation system selected in
Table 8.13. Bearing deformations and force demands
Bearing

Direction

M1a
M1a
M1a
M1a
M1b
M1b
M1b
M1b
A1a
A1a
A1a
A1a
A1b
A1b
A1b
A1b

X
X
Y
Y
X
X
Y
Y
X
X
Y
Y
X
X
Y
Y

Max
Min
Max
Min
Max
Min
Max
Min
Max
Min
Max
Min
Max
Min
Max
Min

u1: m

u2: m

u3: m

u1: rad

u2: rad

u3: rad

N: kN

V2: kN

V3: kN

0.007
0.007
0.006
0.008
0.007
0.007
0.006
0.008
0.002
0.002
0.001
0.002
0.002
0.002
0.001
0.002

0.001
0.001
0.000
0.000
0.001
0.001
0.000
0.000
0.396
0.380
0.126
0.111
0.396
0.380
0.126
0.111

0.003
0.003
0.009
0.010
0.000
0.000
0.001
0.001
0.000
0.000
0.001
0.001
0.000
0.000
0.001
0.001

0.001
0.001
0.003
0.003
0.001
0.001
0.003
0.003
0.001
0.001
0.005
0.005
0.001
0.001
0.005
0.005

0.000
0.001
0.000
0.001
0.001
0.000
0.003
0.002
0.001
0.001
0.002
0.002
0.001
0.001
0.002
0.002

0.015
0.014
0.005
0.004
0.015
0.014
0.005
0.004
0.003
0.002
0.003
0.002
0.003
0.002
0.003
0.002

6659
7271
6273
7657
6659
7271
6274
7657
1532
1974
1096
2409
1531
1974
1096
2409

674
669
375
370
674
669
375
370
0
0
0
0
0
0
0
0

0
0
0
0
159
159
530
529
174
187
595
609
187
174
609
596

221

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.39. Bridge elevation


C0

P1

P2

60.00 m

Triple FPS

10 m

Triple FPS

60.00 m

80.00 m

10 m

Triple FPS

C3

2.5 m

Triple FPS

2.5 m

this case consists of triple friction pendulum bearings (see Section 7.4.3.3). The analysis of the
seismic isolation system is carried out with both the fundamental mode method and nonlinear
time-history analysis. The results of the two analysis methods are compared.
8.4.2
Bridge conguration Design concept
8.4.2.1 Bridge layout
The bridge has a composite steelconcrete continuous deck, with spans of 60 80 60 m and
two solid rectangular 10.0 m tall piers. The lower 8.0 m of the pier has rectangular crosssection 5.0 m by 2.5 m. The seismic isolation bearings are supported on a widened pier head
with rectangular plan 9.0 m  2.5 m and 2.0 m depth. The pier concrete class is C35/45.
Figure 8.39 shows the elevation, and Figure 8.40 the typical deck cross-section of the bridge.
Figure 8.41 presents the layout of the seismic isolation bearings, and Figure 8.42 that of the piers.
The large stiffness of the squat piers, in combination with the high design ground acceleration on
rock (agR 0.40g) leads to the selection of seismic isolation. This selection offers the following
advantages:
g
g
g

a large reduction of constraints due to imposed deck deformation


practically equal and therefore minimised seismic action effects on the two piers (this
would have been achieved even if the piers had unequal heights)
drastic reduction in the seismic forces.

The additional damping offered by the isolators keeps the displacements to a cost-effective level.
8.4.2.2 Seismic isolation system
The seismic isolation system consists of eight sliding bearings with a spherical sliding surface, of
the triple friction pendulum system (triple FPS) type. Two triple FPS bearings support the deck
at each abutment, C0 and C3, or pier, P1 and P2. The triple FPS bearings allow displacements in
Figure 8.40. Deck section

7000

222

Girder No. 2

1100 600 1100

2800

Girder No. 1

Axle of the bridge

12 000

Chapter 8. Seismic design examples

Figure 8.41. Layout of seismic isolation bearings


Y
C0_L

P1_L

P2_L

C3_L
X

C0_R

P1_R

P2_R

C3_R

both the longitudinal and transverse directions with a nonlinear frictional force displacement
relation. The approximate bearing dimensions are:
g
g

at the piers: 1.20 m  1.20 m in plan and 0.40 m in height


at the abutments: 0.90 m  0.90 m in plan and 0.40 m in height.

Figure 8.42. Layout of the piers


Longitudinal section at piers:

C
B

B
9.00

A
5.00

Pier cross-section BB:


2.50

Pier cross-section AA:


2.50

A
8.00

8.00

2.00

10.00

10.00

2.00

Transverse section at piers:

9.00

5.00

2.50

Pier head plan view CC:


7.00

9.00

223

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 8.14. Reactions at supports due to permanent loads (in MN, both girders)
Self-weight
after
construction

Minimum
equipment
load

Maximum
equipment
load

Total with
minimum
equipment

Total with
maximum
equipment

Time variation
due to creep
and shrinkage

C0
P1
P2
C3

2.328
10.380
10.258
2.377

0.664
2.440
2.441
0.664

1.020
3.744
3.745
1.019

2.993
12.819
12.699
3.041

3.348
14.123
14.003
3.396

0.172
0.206
0.091
0.126

Sum of
reactions

25.343

6.209

9.528

31.552

34.871

0.000

The layout and the labels for the seismic isolation bearings are depicted in Figure 8.41 (X is the
longitudinal direction and Y is the transverse direction). Figure 7.12 shows a typical triple FPS
bearing, and Figure 7.14 its typical forcedisplacement relationship.
The nominal properties of the selected triple FPS bearings for seismic analysis are:
g
g
g

the effective dynamic friction coefcient: md 0.061 (+16% variability of the nominal
value)
the effective radius of the sliding surface: Rb 1.83 m
the effective yield displacement: Dy 0.005 m.

8.4.3
Design for horizontal non-seismic actions
8.4.3.1 Imposed horizontal loads braking force
Table 8.14 gives the distribution of the reactions on the supports due to permanent loads
according to the gravity load analysis of the bridge. The time variation of permanent reactions
due to creep and shrinkage is very small. So, the reactions due to permanent loads are considered
constant in time.
The minimum value of the longitudinal force at sliding of the whole deck on the bearings is
calculated from the minimum deck weight and the minimum coefcient of friction at the
bearings as
Fy,min 25 500  0.051  1300 kN
This force is not exceeded by the braking load of Fbr 900 kN. Therefore, the pier bearings do
not slide due to braking. As the horizontal stiffness of the abutments is very high, sliding will
occur at the abutments, associated with the development of friction reactions mWa , where Wa
is the corresponding reaction due to permanent loads. The appropriate static system for this
loading therefore has a pinned connection between the pier tops and the deck and a sliding
connection over the abutments with the above friction reactions (Figure 8.43). The total forces
at the abutments may be calculated from the corresponding displacement of the deck and the
forcedisplacement relation of the bearings (additional elastic reaction Wa/R (Figure 8.43)). A
similar situation occurs for the transverse wind loading.

Figure 8.43. Structural system for imposed horizontal loads


*Wa
*Wa
Sliding
Sliding
Pinned
connection

224

Pinned
connection

Chapter 8. Seismic design examples

Figure 8.44. Structural system for imposed deformations

Sliding

Sliding
Pinned
connection

Elastic connection Kpb = Wp/R

Sliding

Friction forces *Wp

8.4.3.2 Imposed deformations that can cause sliding of the pier bearings
Assuming the structural system in the longitudinal direction to be the same as above, the imposed
deformation that can cause sliding at the pier bearings is calculated from the minimum sliding
load of the bearings and the stiffness of the piers:
g

minimum sliding load:


Fy,min 0.051  12 699 648 kN

pier stiffness:
Kpier 3EI/h3 3  34 000 000  [9  (2.5 m)3/12]/(10)3 1 195 313 kN/m

minimum displacement of deck at the pier top to cause sliding:


dmin Fy,min/K 648/1 195 313 0.5 mm

This displacement is very small and is practically exceeded even by small temperature-imposed
deformations. Consequently, sliding occurs at the bearings of at least one of the piers, under
deformations induced by temperature.
8.4.4
Imposed deformation due to temperature variation
A conservative approach for estimating forces and displacements for this case is the following:
because of the inevitable difference between the sliding friction coefcients of the bearings of
the two piers, albeit small, one of the two pier supports is assumed not to slide under nonseismic conditions. The calculation of horizontal support reactions and displacements should
therefore be based on two systems, with the deck pinned on one of the two piers alternatively.
On the other moving supports, an elastic connection is introduced between the deck and the
support, with stiffness equal to Kpb Wp/R (see Figure 7.10, R Rb 1.83 m), calculated on
the basis of a value of Wp equal to the corresponding permanent load. At these supports,
friction forces equal to mWp, should also be introduced, where m is either the minimum or the
maximum value of friction, with opposite signs on the deck and the supporting element, and
directions compatible with the corresponding sliding deformation at the support, as shown in
Figure 8.44. Both displacements and forces can be derived from these systems.
8.4.5

Superposition of the effects of the braking load and imposed deck


deformations
The superposition of the effects of the braking load and imposed deformations should be done
with caution, as the two cases correspond in fact to a nonlinear response of the system due to the
225

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

involvement of the friction forces. The application of the braking force on the system on which
imposed deformations are already acting causes, in general, a redistribution of the friction forces
estimated above: namely that the original friction forces, acting on one of the piers and the corresponding abutment, which had the same direction with the braking force, are reversed, starting
from the abutment, where full reversal, amounting to a force of 2mWa , will take place. The
remaining part of the braking force,
Fbr  2mWa 900  2  0.051  2993 595 kN
should be equilibrated mainly by a decrease in the reaction of the relevant pier. This decrease is
associated with a displacement of the deck in the direction of the breaking force, an upper bound
of which can be estimated as
Dd (Fbr  2mWa)/Kpier 595/1 195 313 0.0005 m 0.5 mm
The corresponding upper bounds of the force increase on the reactions of the opposite pier and
abutment amounts to
DdWp/R 0.0005  12 699/1.83 3.5 kN
and
DdWa/R 0.0005  2993/1.83 0.8 kN
respectively. Consequently, for this example both the displacement Dd and the force increases can
be neglected.
A comparison with the forces and displacements resulting from the seismic design situation
shows that the latter are always governing in a bridge with seismic isolation.
8.4.6
Design seismic action
8.4.6.1 Design spectra
The project-dependent parameters dening the horizontal elastic spectrum (see Section 3.1.2.3 of
this Guide) are:
g
g
g
g
g
g

type 1 horizontal elastic response spectrum


no near source effects
importance factor gI 1.0
reference peak ground acceleration for type A ground agR 0.4g
design ground acceleration for type A ground ag gIagR 0.40g
ground type B with soil factor S 1.20, periods TB 0.15 s, TC 0.5 s and TD 2.5 s.

The value of the period TD is particularly important for the safety of bridges with seismic isolation because it affects proportionally the estimated displacement demands. For this reason,
the National Annex to Part 2 of Eurocode 8 may specify a value of TD specically for the
design of bridges with seismic isolation that is more conservative (longer) than the value
ascribed to TD in the National Annex to Part 1 of Eurocode 8 (CEN, 2004b). For this particular
example, the selected value is TD 2.5 s, which is longer than the value TD 2.0 s recommended
in Part 1 of Eurocode 8.
The project-dependent parameters that dene the vertical response spectrum (Section 3.1.2.4)
are:
g
g
g

type 1 vertical elastic response spectrum


ratio of the design ground acceleration in the vertical direction to the design ground
acceleration in the horizontal direction avg/ag 0.9
periods TB 0.05 s, TC 0.15 s and TD 1 s.

The horizontal and vertical design spectra are shown in Figures 8.45 and 8.46, respectively.
226

Chapter 8. Seismic design examples

Figure 8.45. Horizontal elastic response spectrum


1.40

Spectral acceleration: g

1.20

Damping 5%

1.00
0.80
0.60
0.40
0.20
0.00
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Period: s

8.4.6.2 Accelerograms for nonlinear time-history analysis


For the time-history representation of the seismic action, seven ground motion time histories are
used (EQ1 to EQ7), each consisting of a pair of horizontal ground motion time-history components and a vertical ground motion time-history component. Each component is produced
by modifying natural, recorded accelerograms to match the Eurocode 8 elastic spectrum
(semi-articial accelerograms). The modication procedure consists of applying unit impulse
functions that iteratively correct the accelerogram in order to better match the target
spectrum. No scaling of the individual components is required to ensure compatibility with
the Eurocode 8 spectrum, as each component is already compatible with the corresponding
spectrum owing to the applied modication procedure.
Figure 8.47 shows an example. The original record is from the Loma Prieta (CA) earthquake,
Corralitos 000 record (magnitude Ms 7.1, distance to fault 5.1 km, USGS ground type B).
For the produced accelerogram the acceleration, velocity and displacement time histories are
shown. Comparing the initial recorded accelerogram with the nal semi-articial one, it is
concluded that the modication method does not alter signicantly the natural waveform.
The 5%-damped pseudo-acceleration and displacement response spectra of the semi-articial
Figure 8.46. Vertical elastic response spectrum
1.40
Damping 5%

Spectral acceleration: g

1.20
1.00
0.80
0.60
0.40
0.20
0.00
0.0

0.5

1.0

1.5
Period: s

2.0

2.5

3.0

227

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.47. Example of horizontal semi-articial accelerogram produced by modifying natural record

Acceleration: g

Original record:
0.80
0.60
0.40
0.20
0.00
0.20
0.40
0.60

10

15

20

25

30

35

40

45

25

30

35

40

45

25

30

35

40

45

25

30

Time: s

Acceleration: g

Modified record:
0.80
0.60
0.40
0.20
0.00
0.20
0.40
0.60
0

10

15

20

Velocity: m/s

Time: s
0.40
0.20
0.00
0.20
0.40
0.60
0.80
1.00

10

15

20

Displacement: m

Time: s
0.400
0.300
0.200
0.100
0.000
0.100
0.200
0.300

10

15

20

35

40

45

Time: s
1.40

Maximum acceleration: 0.560 g


at time t = 2.580 s

Damping: 5%

Pseudo acceleration: g

1.20
1.00

Maximum velocity: 83.734 cm/s


at time t = 2.545 s

0.80

Maximum displacement: 35.575 cm


at time t = 2.305 s

0.60

Vmax/Amax: 0.153 s

0.40

Acceleration RMS: 0.060 g


Velocity RMS: 9.017 cm/s
Displacement RMS: 16.542 cm

0.20
0.00
0.0

0.5

1.0

1.5

2.0

2.5

3.0
Period: s

3.5

4.0

4.5

5.0

5.5

6.0

0.600

Displacement: m

Damping: 5%

Arias intensity: 2.186 m/s


Characteristic intensity (Ic): 0.092
Specific energy density: 3243.860 cm2/s
Cumulative absolute velocity (CAV):
1035.219 cm/s

0.500

Acceleration spectrum intensity (ASI): 0.482 g s


Velocity spectrum intensity (VSI): 213.207 cm

0.400

Sustained maximum acceleration (SMA): 0.257 g


Sustained maximum velocity (SMV): 27.014 cm/s
Effective design acceleration (EDA): 0.568 g

0.300

A95 parameter: 0.553 g


Predominant period (Tp): 0.380 s
Mean period (Tm): 0.630 s

0.200

0.100

0.000
0.0

0.5

1.0

1.5

2.0

2.5

3.0
Period: s

3.5

4.0

4.5

5.0

5.5

6.0

accelerogram are also compared with the corresponding Eurocode 8 spectra, matching the target
spectrum for the full range of periods shown.
The consistency of the ensemble of ground motions is veried in accordance with Section 3.1.4, as
depicted in Figures 8.48 and 8.49 for horizontal and vertical components, respectively. It is
veried that the selected accelerograms are consistent with the Eurocode 8 spectrum for all
228

Chapter 8. Seismic design examples

Figure 8.48. Verication of consistency in the mean of accelerograms used for the horizontal
components with the Eurocode 8 spectrum
2.00

Average SRSS spectrum of


ensemble of earthquakes
1.3 elastic spectrum

1.80
Spectral acceleration: g

1.60

Damping 5%

1.40
1.20
1.00
0.80
0.60
0.40
0.20
0.00
0.0

0.5

1.0

1.5

2.0

2.5
Period: s

3.0

3.5

4.0

4.5

5.0

periods between 0 and 5 s for the horizontal components or between 0 and 3 s for the vertical
component. Therefore, consistency is established for isolation systems with an effective period
Teff , 5/1.5 3.33 s and a prevailing vertical period TV , 3/1.5 2 s, which are fullled for
the isolation system of this example.
8.4.7
Modelling of the structural system for seismic analysis
8.4.7.1 Structural model
8.4.7.1.1 Bridge model
For the purpose of nonlinear time-history analysis, the bridge is modelled in 3D with computer
code SAP 2000, fully accounting for the geometry and spatial distribution of the stiffness and
mass of the bridge. The superstructure and the substructure of the bridge are modelled with
linear prismatic beam elements with properties in accordance with the actual cross-section of
the element. Masses are lumped at the nodes of the model. Where necessary, kinematic
constraints were applied to establish proper connection of the elements. The effect of the
foundation exibility is ignored, and piers are taken as xed at their base. The model of the
bridge for the time-history analysis is shown in Figure 8.50.

Figure 8.49. Verication of consistency between target spectrum and mean spectrum of accelerograms
used for the vertical component
1.40

Average ensemble spectrum


0.9 elastic spectrum

1.20
Spectral acceleration: g

Damping 5%
1.00
0.80
0.60
0.40
0.20
0.00
0.0

0.5

1.0

1.5
Period: s

2.0

2.5

3.0

229

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.50. Bridge model for timehistory analysis

8.4.7.1.2 Isolator model


The triple FPS bearings are modelled with nonlinear hysteretic friction elements, connecting deck
and pier nodes at the locations of the corresponding bearing. In SAP 2000 the behaviour of the
isolator elements in the horizontal direction follows a coupled frictional law based on the Bouc
Wen model. In the vertical direction the behaviour of the isolators corresponds to stiff support
acting only in compression. The actual vertical load of the bearings at each time instant is taken
into account to establish the forcedisplacement relation of the bearing. The effects of bridge
deformation and vertical seismic action are taken into account in the estimation of vertical
bearing loads.
8.4.7.1.3 Effective pier stiffness
The effective pier stiffness is taken as equal to the uncracked gross section stiffness. Because the
stiffness of the piers is much larger than the effective stiffness of the isolation system, piers may
be considered as rigid without signicant loss of accuracy. This approach is followed in the
fundamental mode analysis, presented later with detailed manual calculations. In the nonlinear
time-history analysis, the effective pier stiffness is included.
8.4.8
Bridge loads for the seismic analysis
8.4.8.1 Permanent loads
Permanent action effects vary little with time due to creep and shrinkage (see Table 8.14).
Because of this small variation, only the action effects after fully developed creep and shrinkage
are considered. According to the results of the analysis, the longitudinal displacements due to
permanent actions are approximately 8 mm at the abutments and 3 mm at the piers, both
towards the centre.
8.4.8.2 Quasi-permanent trafc loads
According to Part 2 of Eurocode 8, for road bridges with severe trafc (i.e. bridges of motorways
and other roads of national importance) the quasi-permanent value c2,1Qk,1 of the trafc action
to be considered in the seismic design situation is calculated from the UDL system of trafc
model LM1, with the value of the combination factor c2,1 0.2. The division of the carriageway
in three notional lanes according to clause 4.2.3 of EN 1991-2:2003 (CEN, 2003a) is shown in
Figure 8.51.
The values of the UDL system of LM1 are calculated with the adjustment factor for UDL
aq 1.0:
g

lane 1:

aqq1,k 3 m  9 kN/m2 27.0 kN/m


230

Chapter 8. Seismic design examples

Figure 8.51. Division of carriageway into notional lanes


Modelled girder
1.00

0.50
3.00

3.00

2.00

Lane No. 2

Lane No. 3

Residual area

Girder No. 1

Axle of the bridge

3.00
Lane No. 1

Girder No. 2

3.50

3.50

lane 2:

aqq2,k 3 m  2.5 kN/m2 7.5 kN/m


g

lane 3:

aqq3,k 3 m  2.5 kN/m2 7.5 kN/m


g

residual area:

aqqr,k 2 m  2.5 kN/m2 5.0 kN/m


Total load 47.0 kN/m.
The quasi-permanent trafc load per unit of length of the bridge in the seismic design situation is:

c2,1Qk,1 0.2  47 kN/m 9.4 kN/m


The reactions of the deck supports for the quasi-permanent trafc load are presented in
Table 8.15, according to the analysis of the bridge for the UDL system of LM1.
8.4.8.3 Total weight on the deck in the seismic design situation
The weight Wd of the deck in the seismic design situation includes the permanent loads and the
quasi-permanent value of the trafc loads:
Wd dead load quasi-permanent trafc load 34 871 1880 36 751 kN
Table 8.15. Total reactions due to the quasi-permanent trafc loads
Reactions due to the quasi-permanent
trafc load (for both girders): MN
C0
P1
P2
C3

0.201
0.739
0.739
0.201

Sum of reactions

1.880

231

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

8.4.8.4 Thermal action


The minimum ambient air temperature with a mean return period of 50 years for the structure
is assumed equal to Tmin 208C. The maximum ambient air temperature again with a mean
return period of 50 years is assumed equal to Tmax 408C. The initial temperature is taken
as T0 108C. The uniform bridge temperature components Te,min and Te,max are calculated
from Tmin and Tmax using Figure 6.1 in EN 1991-1-5:2003 (CEN, 2003b), for a type 2 deck
(i.e. composite). The ranges of the uniform bridge temperature component are calculated as:
g

maximum contraction range:


DTN,con T0  Te,min 108C  (208C 58C) 258C

maximum expansion range:


DTN,exp Te,max  T0 (408C 58C)  108C 358C.

According to note 2 in clause 6.1.3.3(3) of EN 1991-1-5:2003, for the design of bearings and
expansion joints the temperature ranges are increased as follows:
g

maximum contraction range for bearings:


DTN,con 208C 258C 208C 458C

maximum expansion range for bearings:


DTN,exp 208C 358C 208C 558C

8.4.9
Design properties of the isolators
8.4.9.1 Upper- and lower-bound design properties
The nominal values of the design properties of the isolators have been presented in Section
8.4.2.2. They are:
g
g
g

the effective dynamic friction coefcient md 0.061 (+16% variability with respect to the
nominal value)
the effective radius of sliding surface Rb 1.83 m
the effective yield displacement Dy 0.005 m.

As pointed out in Section 7.2.2.4, two sets of design properties of the isolating system are
considered:
g
g

the upper-bound design properties (UBDP)


the lower-bound design properties (LBDP).

A separate analysis is performed for each one. For the selected isolation system, only the effective
dynamic friction coefcient md is subject to variability of its design value. The effective radius of
the sliding surface Rb is a geometric property not subject to variability. The UBDP and the LBDP
for md are calculated according to Annexes J and JJ of Part 2 of Eurocode 8:
g
g
g

Nominal value: md 0.061 + 16% 0.051 to 0.071


LBDP: md,min min DPnom 0.051
UBDP: according to Annexes J and JJ of Part 2 of Eurocode 8.

8.4.9.2

Minimum isolator temperature for seismic design

Tmin,b c2Tmin DT1 0.5  (208C) 5.08C 5.08C


where c2 0.5 is the combination factor for thermal actions for the seismic design situation;
Tmin 208C is the minimum shade air temperature at the bridge location having an annual
probability of exceedance of 0.98, per clause 6.1.3.2 of EN 1991-1-5:2003; and DT1 5.08C
232

Chapter 8. Seismic design examples

is the correction temperature for composite bridge decks according to Table J.1N of Part 2 of
Eurocode 8.
8.4.9.3 lmax factors per Annex JJ of Part 2 of Eurocode 8
f1 ageing: lmax,f1 1.1 (Table JJ.5, for normal environment, unlubricated PTFE and
protective seal).
f2 temperature: lmax,f2 1.15 (Table JJ.6 for Tmin,b 5.08C, unlubricated PTFE).
f3 contamination lmax,f3 1.1 (Table JJ.7 for unlubricated PTFE and sliding surface facing
both upwards and downwards).
f4 cumulative travel lmax,f4 1.0 (Table JJ.8 for unlubricated PTFE and cumulative
travel  1.0 km).
Combination factor c 0.70 for importance class II (Table J.2).
Combination value of lmax factors: lU, 1 (lmax,  1)c (Eq. (J.5)).
f1 ageing: lU,f1 1 (1.1  1)  0.7 1.07.
f2 temperature: lU,f2 1 (1.15  1)  0.7 1.105.
f3 contamination lU,f3 1 (1.1  1)  0.7 1.07.
f4 cumulative travel lU,f4 1 (1.0  1)  0.7 1.0.
8.4.9.4

Effective UBDP

UBDP max DPnomlU,f1lU,f2lU,f3lU,f4

(Eq. (J.4))

md,max 0.071  1.07  1.105  1.07  1.0 0.071  1.265 0.09


Therefore, the range of variation of the effective friction coefcient md is from 0.051 to 0.09.
8.4.10 Analysis with the fundamental mode method
8.4.10.1 General
The fundamental mode method of analysis is described in Section 7.5.3 of this Guide. In each of
the two horizontal directions of the seismic action the response of the isolated bridge is determined considering the superstructure as a linear SDoF system using:
g
g
g
g

the effective stiffness of the isolation system Keff


the effective damping of the isolation system jeff
the mass of the superstructure Md
the spectral acceleration Se(Teff , jeff) corresponding to the effective period Teff and the
effective damping jeff .

The effective stiffness at each support location is the composite stiffness of the isolator unit and
the corresponding substructure, per Eqs (D2.10) or (D7.37). In this particular example, the stiffness of the piers is much larger than the stiffness of the isolators, and the contribution of the pier
stiffness may be ignored (see Eq. (D7.32)). The effective damping is derived from Eq. (D7.33) at
the design displacement dcd . The value of dcd is calculated from the effective period Teff and the
effective damping jeff, both of which depend on the value of the still unknown design displacement, dcd . Therefore, the fundamental mode method is in general an iterative procedure, where a
value is assumed for the design displacement in order to calculate Teff and jeff , and a better
approximation of dcd is then calculated from the design spectrum using Teff and jeff . The new
value of dcd is used as the initial value for the new iteration. The procedure converges rapidly.
In this example, hand calculations are presented for the fundamental mode analysis for both
LDBP and UBDP. Only the rst and the last iteration are presented.
8.4.10.2 Fundamental mode analysis for LBDP
The analysis below corresponds to the LBDP of isolators (i.e. md 0.051). The iteration steps are
presented in detail. It is recalled that the weight is Wd 36 751 kN (see Section 8.4.8.3).
Iteration 1
Assume a value for the design displacement dcd: assume dcd 0.15 m.
Effective stiffness of the isolation system Keff (ignoring the piers):
Keff F/dcd Wd[md dcd/Rb]/dcd 36 751  [0.051 0.15/1.83]/0.15 32 578 kN/m
233

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Effective period of the isolation system Teff (Eq. (D7.34)):


Teff

r
r
m
36 751=9:81
2:13 s
2p
2p
Keff
32 578

Dissipated energy per cycle ED:


ED 4Wdmd(dcd  Dy) 4  36 751  0.051  (0.15  0.005) 1087.09 kN m
Effective damping jeff:

jeff

ED,i/(2pKeffd2cd) 1087.09/2p  32 578(0.15)2] 0.36

p
heff [0.1/(0.05 jeff)] 0.591

Design displacement dcd (Table 8.1 in Part 2 of Eurocode 8):


dcd (0.625/p2)agSheffTeffTC (0.625/p2) (0.4  9.81)  1.2  0.591
 2.13  0.5 0.188 m
Assumed displacement: 0.15 m. Calculated displacement: 0.188 m ) second iteration.
Iteration 2
Assume a new value for the design displacement dcd: assume dcd 0.22 m.
Effective stiffness of the isolation system Keff:
Keff 36 751  (0.051 0.22/1.83)/0.22 m 28 602 kN/m
Effective period of the isolation system Teff:
Teff

r
r
m
36 751=9:81
2p
2:27 s
2p
Keff
28 602

Dissipated energy per cycle ED:


ED 4  36 751  0.051  (0.22  0.005) 1611.90 kN m
Effective damping jeff:

jeff 1611.9/[2p 28 602  (0.22)2] 0.1853


p
heff [0.1/(0.05 jeff)] 0.652
Design displacement dcd:
dcd (0.625/p2)  0.4  9.81  1.2  0.652  2.27  0.5 0.22 m
Assumed displacement calculated displacemen ) convergence.
Spectral acceleration Sa:
Sa 2.5(TC/Teff)heff
agS 2.5  (0.5/2.27)  0.652  0.4  1.2 0.172g
Isolation system shear force Vd:
Vd Keffdcd 28 602  0.22 6292 kN
234

Chapter 8. Seismic design examples

8.4.10.3 Fundamental mode analysis for UBDP


For the UBDP of isolators: md 0.09.
Iteration 1
Assume a value for the design displacement dcd: assume dcd 0.15 m.
Effective stiffness of the isolation system Keff:
Keff 36 751  (0.09 0.15/1.83)/0.15 42 133 kN/m
Effective period of the isolation system Teff:
Teff

r
r
m
36 751=9:81
1:87 s
2p
2p
Keff
42 133

Dissipated energy per cycle ED:


ED 4  36 751  0.09  (0.15  0.005) 1984.55 kN m
Effective damping jeff:

jeff 1984.55/[2p  42 133  (0.15)2] 0.333


p
heff [0.1/(0.05 jeff)] 0.511
Calculate design displacement dcd:
dcd (0.625/p2)  0.4  9.81  1.2  0.511  1.87  0.5 0.142 m
Assumed displacement: 0.15 m. Calculated displacement: 0.142 m ) do another iteration.
Iteration 2
Assume a new value for the design displacement dcd: assume dcd 0.14 m.
Effective stiffness of the isolation system Keff:
Keff 36 751  (0.09 0.14 m/1.83 m)/0.14 43 541 kN/m
Effective period of the isolation system Teff:
Teff

s
r
m
36 751kN=9:81m=s2 
1:84 s
2p
2p
Keff
43 541 kN=m

Dissipated energy per cycle ED:


ED 4  36 751  0.09  (0.14  0.005) 1799.32 kN m
Effective damping jeff:

jeff 1799.32/[2p  43 541  (0.14 m)2] 0.331


p
heff [0.1/(0.05 jeff)] 0.512
Calculate design displacement dcd:
dcd (0.625/p2)  0.4  9.81  1.2  0.512  1.84  0.5 0.14 m
Assumed displacement: 0.14 m. Calculated displacement: 0.14 m ) convergence achieved.
Spectral acceleration Sa:
Sa 2.5  (0.5/1.84)  0.512  0.4  1.2 0.166g
235

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Isolation system shear force Vd:


Vd 43 541  0.14 6096 kN
Typically, LBDP analysis leads to maximum displacements of the isolating system, and UBDP
analysis leads to maximum forces in the substructure and the deck. However, the latter is not
always true, as demonstrated by this example. In this particular example, the LBDP analysis
leads to larger shear force (Vd 6292 kN) in the substructure than the UBDP analysis
(Vd 6096 kN). This is because the increase in forces due to the reduced effective damping in
the LBDP analysis (jeff 0.1853 for LBDP versus jeff 0.331 for UBDP) is more dominant
than the reduction in forces due to the increased effective period in the LBDP analysis
(Teff 2.27 s in LBDP versus Teff 1.84 s in UBDP).
8.4.11 Nonlinear time-history analysis
8.4.11.1 Analysis method
The nonlinear time-history analysis for the ground motions of the design seismic action is
performed with direct time integration of the equation of motion using the Newmark constant
acceleration integration scheme with parameters g 0.5 and b 0.25. The integration time
step is taken equal to 0.01 s, which is then divided to the half value if convergence is not
achieved. At each iteration, convergence is considered to be achieved when the unbalanced
nonlinear force is less than 10  4 of the total force.
The damping matrix C is determined from Eq. (D7.29) (Rayleigh damping) with the coefcient
values in Eq. (D7.31). Figure 8.52 shows the damping ratio corresponding to the applied
damping matrix C as a function of mode period. The damping for periods T . 1.5 s, where
seismic isolation dominates, is very small (j , 0.3%). For that period range, energy dissipation
occurs primarily from the nonlinear response of the isolators. For very short periods (T , 0.05 s),
damping increases signicantly (j . 10%). This is desirable, because modes with periods of the
same order of magnitude as the time step cannot be integrated with good accuracy, and it is
preferable to lter them out via increased damping.
8.4.11.2 Action effects on the seismic isolation system
Figures 8.538.56 depict the hysteresis loops for an abutment bearing (C0_L) and a pier bearing
(P1_L) for both the LBDP and UBDP analyses. In Tables 8.16 and 8.17 the time-history analysis
results are presented for the left and right bearings at each pier (P1_L, P1_R, P2_L and P2_R)
and abutment location (C0_L, C0_R, C3_L and C3_R). As the analysis is carried out for
Figure 8.52. Damping as a function of the period of the modes
20.0
17.5

Damping, : %

15.0
12.5
10.0
7.5
5.0
2.5
0.0
0.0

236

0.5

1.0

1.5

2.0
Period, T : s

2.5

3.0

3.5

4.0

Chapter 8. Seismic design examples

Figure 8.53. Hysteresis loops for abutment bearing C0_L from the analysis with LBDP

Direction X

0.200

400
300
200
100
0
0.100 100 0
200
300
400

Direction Y
600
400

0.100

0.200

0.300

Force: KN

Force: KN

EQ1

200
0.300

0.200

0
0.100
0
200

600

400

400
Force: KN

Force: KN

600

0
0.200

0.100

200

0.100

0.200

0.300

0.200

0.100

400

0.100

0.200

Force: KN

Force: KN

0.100
200

200

0.300

0.200

0
0.100
0
200

400

400
300
200
100
0
0.100 100 0
200
300
400

100
0.050 0.100 0.150 0.200

200

Force: KN

Force: KN

200

0.300

0.200

300
400
Displacement: m

400

400

300

0
0

0.100

0.200

Force: KN

Force: KN

600

200

0.200

0.100

600

300

300

600

200

400

100

0.050

0.100

0.150

0.100

200
300

400

200

300

0.100

0.200

Force: KN

Force: KN

200

0.200

400

200
0
100 0

Displacement: m

300

0.100

0.100

0
0.200

200

100
0.200

0.200

200

Displacement: m

EQ7

0.100

Displacement: m

Force: KN

Force: KN

0.050

0
100 0
200

0
0.100

0.200

100

400

100
0.150

0.100

200

Displacement: m

EQ6

0.300

Displacement: m

200
0.100

0.200

Displacement: m

300

0
0.200 0.150 0.100 0.050
100 0

0.100

400

Displacement: m

0.200

0.100

600

0.300

Displacement: m

200

EQ5

200
400

400

EQ3

EQ4

0.300

0
0.300

Displacement: m

0.200

0.200

200

400

0.300

0.300

Displacement: m

200

0.300

0.200

400

Displacement: m

EQ2

0.100

0.300

0.200

0
0.100
0
200

0.100

0.200

0.300

400

400

600

Displacement: m

Displacement: m

237

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.54. Hysteresis loops for abutment bearing C0_L from the analysis with UBDP

Direction X

Direction Y
400
300
200
100
0
0.050100 0
200
300
400

400

EQ1

0.200

0.150

0
0.050
0
200

0.100

0.050

0.100

0.150

Force: KN

Force: KN

200

0.200

0.150

0.100

400

Displacement: m

0.200

0.100

500
400
300
200
100
0
100 0
200
300

400

0.100

0.200

0.300

200
0
0.150 0.100 0.050
0
200

600

200

400

0.050
0
200

0.050

0.100

0.150

Force: KN

Force: KN

400

400

0
0.150 0.100 0.050
0
200

600

0.050

0.100

0.150

Force: KN

Force: KN

0
100 0

0.150

0.100

200
300

Displacement: m

600

400

400

300

200
0
0.150

0.100

0.050
0
200

0.050

0.100

0.050

0.100

0.050

600

300

0.050

0.100

0.150

0.100

400
300
200
100
0
0.050100 0
200
300
400

400

100

300

Displacement: m

0.050

0.100

Force: KN

Force: KN

600

200

200

238

0.050

300

0
0.050

0.050

0.100

0.100

0.150

Displacement: m

100
0.100

0.100

Displacement: m

Displacement: m

0.150

0.050

0
100 0
200

400
300
200
100
0
100 0
200
300
400

EQ7

0.200

100

400

Force: KN

Force: KN

0.100

0.150

200

Displacement: m

EQ6

0.100

Displacement: m

Force: KN

Force: KN

EQ5

0.200

0.050

400
300
200
100
0
0.050 100 0
200
300
400

300
100

0.250

0.250

Displacement: m

400

0.050

0.200

400

200

0.100

0.150

200

Displacement: m

EQ4

0.100

Displacement: m

0
0.100

0.050

400

EQ3

0.150

0.150

600

Displacement: m

0.200

0.100

Displacement: m

Force: KN

Force: KN

EQ2

0.050

200

0.200

0.150

0.100

0
0.050
0
200
400

Displacement: m

0.050

0.100

0.150

Chapter 8. Seismic design examples

Figure 8.55. Hysteresis loops for pier bearing P1_L from the analysis with LBDP

Direction X

0.200

2000

1000

1500

0
0.100
0.000
500

0.100

0.200

0.300

Force: KN

Force: KN

EQ1

Direction Y

1500

1000
500
0.300

0
0.100 5000.000

0.200

1000
1500

2000

2000

1500

1500

1000

1000

0
0.100 5000.000

0.100

0.200

0.300

Force: KN

Force: KN

0.200

500
0.300

0
0.100 5000.000

0.200

1000

1500

Displacement: m

Displacement: m
1500

1000

1000

0.100

0
5000.000

0.100

0.200

1000

Force: KN

Force: KN

1500
500
0.200

0.200

0.100

1500
2000

1500

1500

500

1000
Force: KN

Force: KN

1000

0.300

0.200

0.100

1000

0.100

0.200

500
0.150

0.100

0
0.050
0.000
500

1000

0.050

0.100

Force: KN

Force: KN

1500

500

0.200

0.100

0.100

0
0.000
500
1000

1500

1500

0.100

1500

1500

1000

1000

500

500

1000
1500
Displacement: m

0.200

Displacement: m

0.100

0.200

Force: KN

Force: KN

0.200

0.150

500

1000

0
0.000
500

0.100

Displacement: m

1000

0
0.000
500

0.050

1000

Displacement: m

EQ7

0.200

1500

Force: KN

Force: KN

2000
1500
1000
500
0
0.100 5000.000
1000
1500
2000

0.050

0.100

Displacement: m

Displacement: m

0.100

0
0.000
500

1500

Displacement: m

0.150

0.200

1000

1500

EQ6

0.100

500

1000

0.200

0.300

Displacement: m

0.200 0.150 0.100 0.050 0.000 0.050 0.100 0.150 0.200


500

0.300

0
0.000
500
1000

EQ5

0.200

500

Displacement: m

EQ4

0.100

1000

1500

EQ3

0.300

0.300

Displacement: m

500
0.300

0.200

1500
Displacement: m

EQ2

0.100

1000

0.300

0.200

0.100

0
0.000
500

0.100

0.200

1000
1500
Displacement: m

239

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.56. Hysteresis loops for pier bearing P1_L from the analysis with UBDP

Direction X
EQ1

Direction Y
1500

1500

1000

500
0
0.050500
0.000

0.200 0.150 0.100

0.050

0.100

0.150

Force: KN

Force: KN

1000

1000

500

0.150

0.100

0.050

1500

2000

1500

1500

500
0.100

0
5000.000

0.100

0.200

0.300

0.150 0.100

0
0.050 0.000
500

1500

2000

1000

1500

0.050

0.100

0.150

Force: KN

Force: KN

1500
500

1000
1500

0.050

0
0.150 0.100 0.050 0.000
500

0.050

1500
1000

500

500

0.100

0.150

0.150

0.100

0.050

1500

1500

1000

1000

500
0
0.250 0.200 0.150 0.100 0.0505000.000

0.050

0.100

Force: KN

Force: KN

EQ5

0.100

0.050

1500

1000

1000

500

500

0.050

0.100

Force: KN

Force: KN

1500

0
0.000
500

0.100

0.050

0
0.000
500

1000

1500

500

1000

0
0.000
500
1000

1500
Displacement: m

0.050

0.100

Force: KN

Force: KN

0.150

0.050

0.100

1500
Displacement: m

1500
Displacement: m

0.050

0.100

1000

1000

EQ7

0.050

1500
Displacement: m

2000
Displacement: m

240

0
0.000
500
1000

1500

0.100

0.050

500

1000

0.150

0
0.000
500

1500
Displacement: m

Displacement: m

0.050

0.200

1000

1500

EQ6

0.150

Displacement: m

1000

0.100

0.100

1000

0.050

0.200

500

1500

0
0.000
500

0.150

1000

Force: KN

Force: KN

0.100

0.100

1000

2000
Displacement: m

EQ4

0.050

1000
Displacement: m

Displacement: m

0
0.200 0.150 0.100 0.0505000.000

0.150

500

1000

EQ3

0.100

1000

Force: KN

Force: KN

2000
1000

0.200

0.050

1000
Displacement: m

2000
Displacement: m

EQ2

0
0.000
500

500
0
0.200 0.150 0.100 0.050 0.000
500
1000
1500
Displacement: m

0.050

0.100

0.150

Chapter 8. Seismic design examples

Table 8.16. Bearings: results of the analysis for LBDP


Bearing

dEd,x: m

dEd,y: m

dEd: m

aEd: rad

NEd,min: kN

C0_L
C0_R

0.193
0.193

0.207
0.207

0.255
0.254

0.00498
0.00509

848.7
860.4

C3_L
C3_R

0.199
0.199

0.207
0.207

0.258
0.257

0.00486
0.00494

P1_L
P1_R

0.188
0.188

0.193
0.192

0.244
0.243

P2_L
P2_R

0.189
0.189

0.193
0.192

0.245
0.243

NEd,max: kN

VEd,x: kN

VEd,y: kN

VEd: kN

3310.3
3359.4

346.0
363.2

375.7
389.8

469.0
482.4

855.3
858.5

3323.9
3309.3

402.5
418.4

372.0
368.4

501.4
496.0

0.00367
0.00381

4541.1
4435.4

12086.0
11994.8

1328.5
1369.8

1295.0
1284.5

1654.2
1690.0

0.00369
0.00380

4560.3
4498.0

12084.6
11912.9

1336.1
1365.0

1283.5
1283.2

1654.3
1688.5

6929.3

6652.1

Total

seven seismic motions EQ1 to EQ7, the average of the individual responses may be assumed as
the design value.
The results are the combined effect of the seismic action and the quasi-permanent loads. They do
not include the effects of temperature and creep or shrinkage. dEd,x denotes the displacement in
the longitudinal direction, dEd,y that in the transverse, dEd is the magnitude of the displacement
vector in the horizontal plane and aEd is the magnitude of the rotation vector in the horizontal
plane. NEd is the vertical force on the bearing (positive when compressive), VEd,x is the horizontal
force of the bearing in the longitudinal direction, VEd,y is that in the transverse direction and VEd
is the magnitude of the horizontal force vector.
8.4.11.3 Check of the lower bound on action effects
According to clauses 7.5.6(1) and 7.5.5(6) of Part 2 of Eurocode 8, the resulting displacement of
the stiffness centre of the isolating system (dcd) and the resulting total shear force transferred
through the isolation interface (Vd) in each of the two-horizontal directions are subject to
lower bounds equal to 80% of the design displacement and the shear force through the isolation
interface from the fundamental mode analysis, dcf and Vf, respectively. The lower bounds apply
for both the modal response spectrum analysis and the time-history analysis. The verication of
these bounds is presented below:
g
g
g
g

displacement in the X direction: rd dcd/df 0.193/0.22 0.88 . 0.80 ) bound met


displacement in the Y direction: rd dcd/df 0.207/0.22 0.94 . 0.80 ) bound met
total shear in the X direction: rv Vd/Vf 6929.3/6292 1.10 . 0.80 ) bound met
total shear in the Y direction: rv Vd/Vf 6652.1/6292 1.06 . 0.80 ) bound met.

Witness that the time-history analysis results are 12% smaller for displacements and 10% larger
for the total shear force compared with those of the fundamental mode analysis. This discrepancy
Table 8.17. Bearings: results of the analysis for UBDP
Bearing

dEd,x: m

dEd,y: m

dEd: m

aEd: rad

C0_L
C0_R

0.149
0.149

0.139
0.139

0.182
0.181

0.00469
0.00475

655.0
624.1

C3_L
C3_R

0.157
0.157

0.139
0.138

0.185
0.185

0.00466
0.00461

P1_L
P1_R

0.149
0.149

0.128
0.128

0.173
0.172

P2_L
P2_R

0.150
0.149

0.128
0.127

0.173
0.173

Total

NEd,min: kN

NEd,max: kN

VEd,x: kN

VEd,y: kN

VEd: kN

3157.9
3110.3

352.6
363.4

380.4
366.8

449.8
452.3

677.2
684.8

3112.5
3096.8

400.6
390.6

368.6
360.1

489.6
473.0

0.00361
0.00355

3912.7
3781.8

11246.7
11408.5

1361.8
1352.6

1273.8
1185.7

1630.8
1587.1

0.00359
0.00354

3793.6
3886.4

11246.2
11378.4

1379.7
1370.1

1255.4
1187.1

1605.7
1603.4

6971.3

6377.8

241

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

between the comparison of displacements and forces is attributed to the effect of the vertical
earthquake component on bearing forces, which is not taken into account in the fundamental
mode method. For spherical sliding bearings, the horizontal bearing shear forces are always
proportional to the vertical bearing loads. The variation in the vertical bearing loads due to
the vertical ground motion also affects the horizontal shear forces. This effect is evident in the
wavy nature of the forcedisplacement hysteresis loops of the isolators presented in
Figures 8.538.56.
8.4.12 Verication of the isolation system
8.4.12.1 Displacement demands on the isolation system
The displacement demand in each direction dm,i is determined as the sum of:
g
g

the seismic design displacement, dbi,d , multiplied by an amplication factor gIS with the
recommended value gIS 1.50
the offset displacement dG,i due to quasi-permanent actions, long-term deformations and
50% of the thermal action (cf. Eq. (D6.36) in Section 6.8.1.2).

The offset displacement due to 50% of the thermal action is determined as follows. The design
value of the uniform component of the thermal action is in the range 258C to 358C.
Assuming that the xed point of thermal expansion/contraction is located at one of the two
piers, this leads to an effective expansion/contraction length LT of 140 m for the abutment
bearings and 80 m for the pier bearings. With the sign corresponding to deck movement
towards the abutment and  to movement towards the bridge centre, the offset displacement
due to 50% of thermal action is:
At abutments:
At the piers:

0.5LT
0.5LT
0.5LT
0.5LT

aDT 0.5  140 000  1.0  10  5  (45) 31.5 mm


aDT 0.5  140 000  1.0  10  5  (55) 38.5 mm
aDT 0.5  80 000  1.0  10  5  (45) 18 mm
aDT 0.5  80 000  1.0  10  5  (55) 22.0 mm

The total offset displacement, including the effects of quasi-permanent actions, long-term
deformations and 50% of the thermal action, is calculated as follows:
At abutments:
At the piers:

Towards
Towards
Towards
Towards

the bridge centre: 8  31.5 39.5 mm


the abutment: 38.5 mm
the bridge centre: 3  18 21 mm
the abutments: 22 mm

According to Part 2 of Eurocode 8, the displacement demand is estimated and checked in the
principal directions and not in the most critical direction. However, this is not adequate for
bearings with the same displacement capacity in all horizontal directions, such as the FPS
bearings. The maximum displacement of the isolator occurs in a direction that does not
coincide in general with one of the two principal directions. The maximum required displacement
demand in the most critical direction may be estimated by examining the time history of the
magnitude of the resultant displacement vector in the horizontal plane XY, including the
effect of offset displacements due to quasi-permanent actions, long-term displacements and
50% of the thermal action.
In Table 8.18 the displacement demand at the abutment and pier bearings is estimated in both
principal directions, alongside the critical displacement demand in the horizontal XY plane,
which in the present case is larger by about 25%. Therefore, the displacement demand of the
isolators is 407 mm for abutment bearings and 382 mm for pier bearings.
8.4.12.2. Restoring capability of the isolation system
The lateral restoring capability of the isolation system is veried per clause 7.7.1 of Part 2 of
Eurocode 8. The equivalent bilinear model of the isolation system is shown in Figure 8.57,
where F0 mdNEd is the force at zero displacement; Kp NEd/Rb is the post-elastic stiffness;
and d0 is the maximum residual displacement for which the isolation system can be in static
242

Chapter 8. Seismic design examples

Table 8.18. Displacement demand on isolators


Displacement demand

For abutments:
C0_L, C0_R, C3_L, C3_R

For piers:
P1_L, P1_R, P2_L, P2_R

In longitudinal direction X
In transverse direction Y
In horizontal plane XY

329
311
407

305
290
382

Maximum

407

382

equilibrium in the considered direction. For an isolation system consisting of spherical sliding
isolators the displacement d0 is:
d0 F0/Kp mdNEd/(NEd/Rb) mdRb
According to clause 7.7.1(2) in Part 2 of Eurocode 8, an isolation system has adequate selfrestoring capability if dcd/d0 . d in both principal directions, where d is a coefcient with the
recommended value d 0.5. This criterion is veried for both UBDP and LBDP of the isolators.
Lower values of the design displacement dcd give results that are more on the safe side:
g

longitudinal direction, LBDP:


dcd/d0 0.193/(0.051  1.83) 2.07 . 0.5

transverse direction, LBDP:


dcd/d0 0.207/(0.051  1.83) 2.22 . 0.5

longitudinal direction, UBDP:


dcd/d0 0.149/(0.09  1.83) 0.90 . 0.5

transverse direction, UBDP:


dcd/d0 0.138/(0.09  1.83) 0.84 . 0.5

Therefore, the restoring capability of the isolation system is adequate without additional increase
in the displacement capacity dm . It is noted that UBDP give more unfavourable results because
dcd is larger and d0 smaller than for LBDP.
8.4.13 Verication of the substructure
8.4.13.1 Action effect envelopes for the piers
In Tables 8.19 and 8.20, action effect envelopes from the time-history analysis (average for the
seven earthquake ground motions EQ1 to EQ7) are given for the substructure. For piers P1
Figure 8.57. Properties of bilinear model for verication of restoring capability of isolator
Force
Kp
F0
Displacement
dcd
d0

d0

243

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Table 8.19. Substructure: envelopes of analysis for LBDP


Envelope

N: kN

VX: kN

VY: kN

T: kN m

MX: kN m

MY: kN m

C0
C0
C0
C0
C0
C0
C0
C0
C0
C0
C0
C0

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

1754.3
6535.1
4930.5
3688.2
5623.1
4124.3
2759.9
2989.8
4930.5
3688.2
3789.3
 4324.0

18.3
347.5
616.5
660.5
617.2
469.5
358.3
341.2
616.5
660.5
383.9
493.4

158.3
123.9
163.2
115.7
684.1
694.6
393.2
505.2
163.2
115.7
608.9
730.7

14.6
23.1
85.1
82.2
192.6
190.5
183.1
216.0
85.1
82.2
272.1
312.4

1.8
34.7
61.6
66.0
61.7
47.0
35.8
34.1
61.6
66.0
38.4
49.3

824.8
380.1
475.4
482.2
1933.0
2002.9
1388.7
1867.7
475.4
482.2
2575.8
2701.2

C3
C3
C3
C3
C3
C3
C3
C3
C3
C3
C3
C3

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

1787.9
6439.8
4241.8
3389.9
5460.4
4149.3
1975.2
2760.7
4241.8
3389.9
4001.7
4533.2

105.4
379.4
783.1
562.1
666.9
401.9
257.9
312.5
783.1
562.1
453.0
591.8

113.9
134.5
110.8
106.0
680.5
660.4
301.0
435.8
110.8
106.0
631.7
690.8

31.5
32.2
56.9
66.8
238.2
172.9
172.4
215.7
56.9
66.8
312.7
395.7

10.5
37.9
78.3
56.2
66.7
40.2
25.8
31.2
78.3
56.2
45.3
59.2

654.5
446.2
328.9
429.4
2046.7
1867.8
1131.7
1809.1
328.9
429.4
2622.4
2597.2

P1
P1
P1
P1
P1
P1
P1
P1
P1
P1
P1
P1

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

12756.8
27232.5
16241.5
17636.3
16658.7
15829.2
8022.6
13056.5
16142.4
18598.0
16393.7
18669.2

50.1
228.2
3339.4
2906.9
1112.7
909.9
961.5
2514.3
3319.0
2499.2
1095.0
1073.1

236.8
640.6
500.1
86.6
2666.1
2698.2
813.0
919.9
497.1
1830.2
2623.7
3182.3

60.2
451.2
105.8
77.7
758.9
450.8
575.0
768.1
105.2
240.9
746.9
531.7

254.0
2143.8
29347.6
22629.6
11127.1
9661.0
9403.4
22613.0
29168.5
26831.4
10950.1
11394.4

3971.0
7982.2
4786.2
1838.1
33869.5
27964.5
12731.0
17367.8
4756.9
15284.0
33330.7
32981.8

P2
P2
P2
P2
P2
P2
P2
P2
P2
P2
P2
P2

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

12560.1
27066.2
16266.7
17867.1
16650.4
15988.2
7732.6
12784.1
16195.0
18734.3
16276.3
18798.5

792.5
230.7
3383.2
2879.8
1099.1
956.8
960.9
2478.8
3368.3
2470.9
1074.4
1125.0

174.2
715.6
506.5
84.6
2678.2
2711.5
781.8
860.4
504.3
1809.2
2618.0
3188.1

161.5
339.1
156.6
83.3
777.1
429.9
575.5
766.8
155.9
255.8
759.6
505.5

6724.3
2180.8
29890.9
22406.6
11054.4
10018.0
9395.1
22343.7
29759.0
26514.7
10806.1
11778.9

4432.2
8957.0
4842.4
1807.9
34062.3
28164.0
12189.7
16575.8
4821.0
15186.2
33297.0
33114.5

and P2 they refer to their base, and for the abutments C0 and C3 to the mid-point between the
bearings (at the bearing level). The envelopes include the effect of permanent actions and the
quasi-permanent value of the trafc loads and the design seismic action. They do not include
the effects of temperature and shrinkage. According to clause 7.6.3(2) of Part 2 of Eurocode 8,
244

Chapter 8. Seismic design examples

Table 8.20. Substructure: envelopes of analysis for UBDP


Envelope

N: kN

VX: kN

VY: kN

T: kN m

MX: kN m

MY: kN m

C0
C0
C0
C0
C0
C0
C0
C0
C0
C0
C0
C0

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

1326.0
6076.0
3620.5
3503.1
3737.8
3996.2
2699.6
3260.5
3620.5
3503.1
3222.0
4111.4

116.0
594.2
627.6
693.8
149.9
375.4
22.2
479.0
627.6
693.8
97.5
219.5

80.6
94.3
93.9
158.1
686.9
640.2
197.0
471.3
93.9
158.1
597.8
555.6

12.6
38.4
53.1
133.9
105.3
176.4
300.3
241.3
53.1
133.9
89.8
199.5

11.6
59.4
62.8
69.4
15.0
37.5
2.2
47.9
62.8
69.4
9.7
21.9

62.0
365.0
347.0
687.2
2696.7
2085.3
149.2
1937.6
347.0
687.2
2655.5
2575.7

C3
C3
C3
C3
C3
C3
C3
C3
C3
C3
C3
C3

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

1417.6
6053.3
4215.2
3079.4
4496.6
3930.0
2417.7
2709.4
4215.2
3079.4
3961.3
4233.5

76.4
614.1
768.4
586.3
636.9
296.3
325.0
390.4
768.4
586.3
570.8
117.5

45.9
86.6
147.3
151.7
669.0
635.4
359.0
425.5
147.3
151.7
622.0
558.0

61.4
37.5
39.3
96.5
347.4
149.4
233.9
285.9
39.3
96.5
418.1
8.8

7.6
61.4
76.8
58.6
63.7
29.6
32.5
39.0
76.8
58.6
57.1
11.8

384.7
339.1
381.3
525.6
2340.9
2069.0
1283.0
1840.4
381.3
525.6
2690.9
2615.1

P1
P1
P1
P1
P1
P1
P1
P1
P1
P1
P1
P1

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

11444.7
25719.8
15188.6
17329.4
16196.9
14597.6
12907.1
11198.9
14829.4
15090.4
15952.1
15337.1

125.6
320.3
3632.5
3190.1
1183.0
1.4
1473.0
2406.9
3546.6
3183.3
1165.1
1.4

566.0
1735.7
168.5
282.4
2666.2
2828.3
804.9
983.3
164.5
127.2
2626.0
2971.5

7.6
382.5
83.2
106.6
949.5
175.7
693.2
1016.0
81.2
99.0
935.1
184.6

1131.0
3219.5
32565.5
27647.3
12871.0
580.6
14798.4
21664.8
31795.4
28584.3
12676.5
610.0

7410.2
22258.8
2160.6
3773.0
33694.2
29913.1
13503.0
18338.9
2109.5
246.6
33185.2
31428.6

P2
P2
P2
P2
P2
P2
P2
P2
P2
P2
P2
P2

max N
min N
max VX
min VX
max VY
min VY
max T
min T
max MX
min MX
max MY
min MY

11479.8
25746.2
15433.8
15216.5
20549.5
14855.2
12267.8
11612.0
15110.2
15128.2
16623.8
15508.7

216.1
28.7
3702.6
3197.1
280.4
49.9
1464.3
2520.1
3625.0
3178.6
340.0
52.1

583.7
1764.3
165.4
106.6
2618.8
2856.8
764.4
953.8
161.9
106.0
2495.7
2982.5

1.8
409.5
75.4
115.6
304.4
190.7
741.6
1006.9
73.8
114.9
120.7
199.1

2007.6
372.2
33190.2
28697.8
3324.9
930.3
14684.7
22796.5
32494.1
28531.2
2377.2
971.2

7643.4
22556.9
2114.8
609.4
29039.5
30281.5
12727.1
17940.3
2070.5
605.9
32509.5
31613.6

the design seismic forces due to the design seismic action alone may be derived from the
time-history analysis forces after division by the q factor for limited ductile/essentially elastic
behaviour, q 1.50. This is not included in the present results, but will be applied at the
design stage of the pier cross-sections.
245

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

The following notation is used:


g
g
g
g

N is the vertical force (i.e. axial force) (positive when acting upwards)
VX is the shear force along the X axis, VY is the shear along the Y axis
T is the torsional moment
MX is the moment in the vertical plane through the X axis (i.e. that produced by an
earthquake in the longitudinal, X, direction), and MY the moment in the vertical plane
through the Y axis (produced by an earthquake acting in the transverse direction, Y).
The signs of VX/MX are the same when their directions are compatible with earthquake
forces acting in the X direction; the signs of VY/MY are the same when their directions are
compatible with earthquake forces acting in the Y direction.

Envelopes of maximum/minimum and concurrent internal forces are presented for each pier/
abutment location. For instance, envelope max N corresponds to the design situation where the
value of the vertical force N is algebraically the maximum. The values of other forces VX, VY,
T, MX and MY at max N envelope are the concurrent forces when N becomes a maximum.
The maximum/minimum and the concurrent forces for each envelope are derived as follows:
The maxima/minima of each force (say max MX , j 17) over all time-steps of the
response history for each motion j 17 are assessed. The design value of the maximum/
minimum of the examined force (say MX,d) is the average of these maxima/minima for the
seven motions.
2 The results of the seismic motion producing the extreme value (say maxmax MX) of these
maxima/minima for all motions and the corresponding time-step are used as the basis for
the assessment of the concurrent values of the other forces. A scaling factor is applied to
these results, equal to the ratio of the design value of the examined force (MX,d) divided by
the extreme value (maxmax MX) (i.e. MX,d/maxmax MX).
1

8.4.13.2 Section verication of the piers


8.4.13.2.1 General
The maximum normalised axial force of the piers is

hk NEd/(Ac fck) 27 232.5/(5  2.5  35 000) 0.062 , 0.08


Therefore, according to clause 6.2.1.1(2) of Part 2 of Eurocode 8, no connement reinforcement
is necessary. However, due to the low axial force the pier should be designed taking into account
the minimum reinforcement requirements for beams and for columns.
8.4.13.2.2 Verification for flexure and axial force
According to clause 7.6.3(2) of Part 2 of Eurocode 8, for the design of the substructures the seismic
forces EE due to the design seismic action alone may be obtained by dividing the analysis forces by
the q factor corresponding to limited-ductile/essentially-elastic behaviour, q  1.50.
Clause 6.5.1 in Part 2 of Eurocode 8 prescribes certain reduced ductility measures (for the connement and restraint of reinforcement buckling). It also offers the option of avoiding these
measures if the piers are designed so that MRd/MEd , 1.3. This option is chosen in this
example, for reasons to become apparent soon. Therefore, for the design of longitudinal
reinforcement the design seismic forces, FEd , are derived from forces derived from the timehistory analysis, FEA , as FEd 1.3FE,A/1.5. For the most adverse combination of so-computed
design seismic action effects, NEd , MX,ed and MY,ed , the required longitudinal reinforcement
amounts to As 21 370 mm2, uniformly distributed all around the section.
8.4.13.2.3 Minimum longitudinal reinforcement
Part 2 of Eurocode 8 does not have a specic requirement for a minimum value of the longitudinal reinforcement ratio. The minimum reinforcement specied in Eurocode 2 for columns
(including those of bridges) is equal to
As,min max(0.1NEd/fyd , 0.002Ac) max[0.1  27 232.5/(500 000/1.15), 0.002  5  2.5]
0.025 m2 25 000 mm2
(i.e. min r 0.2%).
246

Chapter 8. Seismic design examples

Eurocode 2 species for beams a minimum amount of tensile reinforcement to prevent brittle
failure right after exceedance of the cracking moment of the section (i.e. of the tensile concrete
strength). For uniaxial bending, the minimum tensile reinforcement (i.e. of the side of the
section that is in tension) is r1,min max (0.26fctm/fyk , 0.0013), normalised to bd. For concrete
C35/45 with fctm 3.2 MPa and steel grade 500 with fyk 500 MPa, r1,min 0.001664, which
gives for each one of the two long sides of the section
As1,min 0.001664  5000  (2500  80) 20 134 mm2
(i.e. 20 134/5.0 4027 mm2/m of the perimeter). If the same reinforcement density is adopted
also for the two short sides of the section (which prot from the tension reinforcement of the
end of the long sides, too), we obtain 60 400 mm2 for the total minimum reinforcement of the
rectangular section, and rmin 0.00483 0.483%.
In summary:
g
g
g

required longitudinal reinforcement for the ULS design of the section: 21 373 mm2
(r 0.17%)
required minimum longitudinal reinforcement: 60 400 cm2 (rmin 0.483%)
provided longitudinal reinforcement: one layer 128/135 4560 mm2/m or 64 000 mm2 in
total (r 0.51%).

Note that the cross-section of the piers may be substantially reduced.


8.4.13.2.4 ULS verification in shear
According to clause 5.6.2(2) of Part 2 of Eurocode 8, for the design of shear reinforcement,
Eurocode 2 applies with the following additional rules:
g
g

The design action effects are multiplied by the behaviour factor q used in the linear
analysis.
The resistance values, VRd,c , VRd,s and VRd,max derived in accordance with Eurocode 2 are
divided by an additional safety factor gBd1 against brittle failure, with the recommended
value gBd1 1.25. Therefore, the design seismic forces for the ULS design in shear, FEd ,
may be derived from the time-history analysis forces, FEA , as FEd 1.25FE,A.

The shear reinforcement calculated as above is presented below:


g
g
g

required shear reinforcement in the longitudinal direction: 5903 mm2/m.


required shear reinforcement in the transverse direction: 2366 mm2/m.
provided shear reinforcement in the longitudinal direction (116/150 on the perimeter plus
4 112/150 cross-ties with two legs each): 4  2  754 mm2/m 2  1340 mm2/m
8710 mm2/m (rw 0.174%).
provided shear reinforcement in the transverse direction (116/150, only on the perimeter):
2  1340 mm2/m 2680 mm2/m (rw 0.107%).

The provided shear reinforcement satises the minimum requirements of Eurocode 2 for
columns:
g
g

maximum spacing 0.6  min(20  28 mm, 2500 mm, 400 mm) 240 mm; provided
spacing 150 mm
minimum bar diameter max (6 mm, 28 mm/4) 7 mm; provided diameter 12 mm.

The shear reinforcement satises even the minimum with Eurocode 2 requirements for
beams:
g
g

maximum longitudinal spacing sl,max 0.75d 0.75  2420 1815 mm; provided
longitudinal spacing 150 mm
maximum transverse spacing st,max min(0.75d, 600 mm) min(0.75  2420 mm,
600 mm) 600 mm; provided transverse spacing 530 mm
247

Designers Guide to Eurocode 8: Design of Bridges for Earthquake Resistance

Figure 8.58. Layout of pier reinforcement


5.0 m
0.53

0.53

0.53

0.53

0.53

0.53

0.53

0.53

0.53

2.5 m

Stirrups: 4 two-legged 12/15 = 4 2 7.54 cm2/m = 60.3 cm2/m


Perimetric hoop: 1 two-legged 16/15 = 2 13.40 cm2/m = 26.8 cm2/m
Longitudinal reinforcement: 1 layer 28/13.5 = 45.6 cm2/m

p
p
minimum shear reinforcement ratio rw,min 0.08 fck/fyk 0.08 35/500 0.095%;
provided shear reinforcement ratio rw 0.174% in the longitudinal direction and
rw 0.107% in the transverse direction.

The reinforcement at the pier base is shown in Figure 8.58.


8.4.14 Design action effects for the foundation
8.4.14.1 Design actions effects from time-history analysis
Table 8.21 lists action effect envelopes for the design of the foundation, on the basis of the timehistory analysis results. The action effects for the foundation design are derived according to
clauses 7.6.3(4) and 5.8.2(2) of Part 2 of Eurocode 8 for bridges with seismic isolation. The
seismic action effects for the foundation design correspond to the analysis results multiplied
by the q value (q 1.5) used for the design of the substructure (i.e. effectively corresponding
to q 1).
The set of forces that are critical for the foundation design are the maximum/minimum shear
force envelopes for the design of the foundation of the abutment and the maximum/minimum
bending moment at the base of the pier, for the design of the pier foundation. The analysis
results for the seismic design situation are given for piers P1 and P2 at their base and for
abutments C0 and C3 at the midpoint between the bearings (i.e. at the bearing level). These
action effects include those of permanent actions, of the combination value of trafc loads
and of the design seismic action. The signs of the forces for foundation design are as in
Figure 8.37.
8.4.14.2 Comparison with the fundamental mode method
The force and displacement results at the abutments and at the base of the piers from the timehistory analysis are compared here with those of the fundamental mode method. LBDP give the
most unfavourable results with respect to the forces in the substructure.

Table 8.21. Seismic design situation action effects for foundation design

248

Fx: kN

Fy: kN

max Fx envelope
max Fy envelope

783
470

111
695

4242
4124

329
2003

78
47

57
191

max My envelope
max Mx envelope

3625
1095

162
2624

15110
16394

2070
33331

32494
10950

74
747

Location

Envelope

C0, C3
P1, P2

Fz: kN

Mx: kN m

My: kN m

Mz: kN m

Chapter 8. Seismic design examples

Table 8.22. Displacements and total shears at abutment bearings in the longitudinal direction for the
two analysis methods

Time-history analysis
Fundamental mode method

Displacement
demand: mm

Total shear in longitudinal


direction: kN

Total shear in the transverse


direction: kN

407
419

783
683

695
683

Before attempting this comparison, the conclusions of Section 7.5.5.4 on bidirectional excitation
are applied:
g

g
g

Applying the second proposal, the effective value dcd,e of the maximum displacement
(length of the vector resultant of the simultaneous orthogonal displacements) is assumed as
equal to dcd,e  1.15dcd .
The increased value dcd,e should be used also for the estimation of the maximum forces
transferred through the isolator in any direction, as the isolator has no preferred direction.
The vertical seismic motion component also has an effect on the variation of the friction
forces of the isolator. The effect oscillates, with approximately equal positive and negative
values and approximately zero mean value. This effect can be observed in the hysteresis
loops of Figures 8.538.56, for some of the seismic motions used (e.g. EQ7 and EQ3). The
oscillations occur at much shorter periods than those of the horizontal motion,
corresponding to the much higher frequency content of the vertical component (cf. the
elastic spectra of the two components). Consequently, this inuence may be ignored, at
least as far as the maximum displacements are concerned. Regarding the forces, the
application to the friction forces of the 1.15 multiplier estimated above is a convenient
approximation.

The displacement demand of abutment bearings and the total abutment shear are presented in
Table 8.22. The table compares the results of the time-history analysis with those of the fundamental mode method multiplied by 1.15, as explained above. The estimated displacement
demand using the fundamental mode method is 3% larger than that from time-history
analysis. The total shear estimated by the fundamental mode method is 13% less in the longitudinal direction and 3% less in the transverse direction than those from time-history analysis.
REFERENCES

CEN (Comite Europe de Normalisation) (2003a) EN 1991-2:2003: Eurocode 1: Actions on


structures Part 2: Trafc loads on bridges. CEN, Brussels.
CEN (2003b) EN 1991-1-5:2003: Eurocode 1: Actions on structures Part 1-5: General actions
Thermal actions. CEN, Brussels.
CEN (2004a) EN 1992-1-1:2004: Eurocode 2: Design of concrete structures Part 1-1: General
rules and rules for buildings. CEN, Brussels.
CEN (2004b) EN 1998-1:2004: Eurocode 8 Design of structures for earthquake resistance Part 1:
General rules, seismic actions and rules for buildings. CEN, Brussels.
CEN (2005a) BS EN 1992-2:2005: Eurocode 2. Design of concrete structures Part 2: Concrete
bridges Design and detailing rules. CEN, Brussels.
CEN (2005b) EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance Part 2:
Bridges. CEN, Brussels.
Bouassida Y, Bouchon E, Crespo P et al. (2012) Bridge Design to Eurocodes Worked examples.
Worked examples presented at the Workshop Bridge Design to Eurocodes, Vienna, 46 October
2010. JRC European Commission (Athanasopoulou A et al. (eds)).

249

INDEX
Page locators in italics denote figures separate from the corresponding text, locators in bold denote tables.

Index Terms

Links

A
abutments
conceptual design

46

detailing

155159

minimum overlap lengths

135136

modal response spectrum analyses


nonlinear time-history analysis
role

58

5964

81
236

237238

5960

seismic design examples

217

verification

155159

accelerograms

227229

accidental action

89

accidental torsion

99

action effects
detailing

149151

foundations

159160

206207

248249

236

237240

241

217219

236246

248249

67

9297

185191

196197

233234

248249

isolation
lower-bound design properties

241242

pier envelopes

243246

seismic design examples

197200

verification

149151

active faults

29

added mass effects


additional uniform traffic load applications
analyses

8486
195
67118

action components

6869

behaviour factors

6973

design spectra
effective stiffness
fundamental mode

26
100107

isolation

185191

limited ductile piers

211214

215

216

67

68

7392

108

112

197

nonlinear analyses

6768

108109

110117

132135

torsional effects

9899

modal response spectrum analyses

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

annular sections

55

application rules

Arahthos bridge (GR)

42

45

articulated connections

43

4950

artificial time-history records

26

27

assumptions

axial force
seismic design examples

197200

219

246

910

5961

62

159

4753

5556

B
backfills
balanced cantilever methods

7374

bars
buckling prevention

204

rotation capacity

134

221

bearings
analyses

70

capacity design effects


conceptual design

128
4243

45

58

6264

design examples

207208

221

detailing

145155

156158

dimensioning

142155

elastic deformations
imposed deformations
isolation

1314
225
176177

liquefaction

33

modal response spectrum analyses

81

nonlinear time-history analysis

236

180181

241

spherical sliding surfaces

181185

verification

145155

156158

10

6973

behaviour factors
bidirectional excitation

189190

bilinear hysteretic behaviour

174177

blow-count number

22

boundary frames

39

box girder decks

7374

braking force

224

braking loads

225226

bridge models

229

230

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

buckling
longitudinal bars
verification

204

221

146147

buoyant mass effects

86

buried structure flotation

33

C
cables

63

Caltrans Seismic Design Criteria

25

39

62

109110

cantilevers

21

42

45

58

191

242243

7274
capability restorations

173

capacity curves

113

capacity design
conceptual design
detailing

58
123

performance requirements
seismic design examples

124128

1213
201202

204206

206

207
verification

123

cascading collapse

40

cases of low seismicity

16

chord rotations

108109

124128

134135

circular
bearings

144148

columns

53

hoops

133

134

piers

55

103

sections

101

clearance lengths

135140

cohesionless soils
elastic response spectra
stability verifications
strength parameters

22
161
3132

cohesive soils

31

collapse

40

161

columns
capacity design shear
conceptual design
effective torsional rigidity

126128
5359
106

This page has been reformatted by Knovel to provide easier navigation.

136

206

Index Terms

Links

complete quadratic combination (CQC) rule


modal response spectrum analyses

92

nonlinear dynamic analyses

111

nonlinear static analyses

114

complex-valued impedance matrices


compliance criteria
composite pier stiffness

8687
517
188

composite (steelconcrete) decks

44

compression models

39

compression zone confinements

164

conceptual design

3765

abutments

5964

connection choice

4353

decks

3841

ductile piers

109

5964

193194

earthquake resistance

3843

foundations

6465

general rules

3843

isolation

171

piers

174

222224

4445

48

5359

concrete
analyses

72

conceptual design

38

detailing

150

discretisation

7374

effective torsional rigidity


modal response spectrum analyses
verification

79

106
7374

7679

80

150

cone penetration tests (CPTs)


confinement reinforcement

33
203

220221

connections
conceptual design

4353

modal response spectrum analyses

8081

constant shear force

12

continuity slabs see link slabs


continuous decks

correction factors

191

costbenefit analyses
coupling

3841

165166
87

CPTs see cone penetration tests


CQC see complete quadratic combination rule
cracking

80
This page has been reformatted by Knovel to provide easier navigation.

9496

Index Terms

Links

creep

196

critical components

139140

cross-elements

80

curved bridges

138139

cyclic resistance ratios (CRR)

3334

cyclic stress ratios (CSR)

3334

cyclic ultimate chord rotation

134135

136

D
damage limitation seismic action

damped elastic response


nonlinear dynamic analyses
time-history records
damping

111112
2728
1015

conceptual design
detailing

52
142143

145

149155

21

22

23

172

175176

145

149155

3841

4353

5964

detailing

120

150

156159

dimensioning

120

discretisation

7374

elastic response spectra


nonlinear time-history analysis

236

seismic isolation

171

188191

190
soil

32

verification

142143

dead uniform traffic load applications

195

decks
capacity design shear
conceptual design

effective torsional rigidity


elastic response spectra

126128

79

106
21

flexible and nearly straight decks

9596

inflexible and nearly straight decks

9495

linear analyses

9899

modal response spectrum analyses

7384

7579

135137

138

overlap lengths
performance requirements
relative displacements
seismic design examples

9
109110
205

206

225226

231232

stiffness of elements

194

verification

150

156159

This page has been reformatted by Knovel to provide easier navigation.

206

207

Index Terms

Links

decorrelation

28

definitions

degree-of-freedom (DoF) systems


fundamental mode analyses
modal response spectrum analyses

9394

96

74

79

80

26

69

226227

see also single-degree-of-freedom systems


design examples see seismic design examples
design spectra
detailing

119169

bearings

142155

limited ductile behaviour

129

linear analyses

122123

nonlinear analyses

132135

seismic links

140142

diagonal cracking

80

dimensioning
design examples
joint reinforcement
seismic links

201203
130
141142

discretisation

7374

displacement

167

detailing

79

152

elastic response spectra


isolation systems
linear analyses
nonlinear dynamic analyses
seismic isolation
verification

25
242
107110
111
186187
152

DoF see degree-of-freedom systems


double spherical sliding surfaces
dry cohesionless soils

182183
161

ductility
analyses methods

6768

balance with strength

1516

conceptual design

5859

design for
detailing
effective flexural stiffness
global response
modal response spectrum analyses

1015

122123

124128

100
1213
80

nonlinear analyses

110111

seismic design examples

193221

verification

122123

114
132135

This page has been reformatted by Knovel to provide easier navigation.

115

87

Index Terms

Links

Duzce (TR) earthquake

40

dynamic amplification

99

dynamic analyses

108109

111114

E
effective stiffness
analyses

100107

design examples

213214

linear analyses
seismic isolation
effective torsional rigidity

230

107
174175

188

106

elastic analyses
design spectra
elastic behaviour

26

69

12

182

elastic deformations
flexible bearings

1314

foundation ground

1314

elastic multi-frame models

39

elastic nodal inertia loads

96

elastic predictions
displacement calculations

108

elastic response
design ground displacement/velocity
design seismic actions
detailing

25
1925
123

horizontal

2124

near-source effects

25

topographic amplification

25

verification

123

vertical component

2425

elastomeric connections
capacity design effects
conceptual design

128
42

43

5053

5556

detailing

145155

dimensioning

142155

modal response spectrum analyses


verification

81
145155

elastoplastic energy dissipators


encased non-laminated elastomeric bearings

177

179

47

This page has been reformatted by Knovel to provide easier navigation.

45

47

Index Terms

Links

energy dissipation
design

1015

seismic isolation

171

172

see also damping


envelope action effects

243246

envelope linear models


conceptual design

39

equal displacement rule

10

equivalent static analyses


exemptions from Eurocode application

9297
16

F
failure

3132

43

47

4950

fixed connections
conceptual design

5556

58
dimensioning

142

fixings
elastomeric bearings

148149

flat sliding bearings


dimensioning
seismic isolation
flexible bearings

142
180181
1314

flexible deck models


flexible and nearly straight decks
flexural force

70

99
9596
197198

219

246

206

206

flexural plastic hinges


analyses

70

detailing/verification

124126

flexural resistance
verification procedures

122123

flexural rigidity
modal response spectrum analyses

80

flexural stiffness
analyses

100101

flexural verification
seismic design examples

205

flotation
liquefaction

33

flow failures

32

FMM see fundamental mode method of analysis


footing structural design

162163

This page has been reformatted by Knovel to provide easier navigation.

207

Index Terms

Links

force reduction factors


force-based approaches

10

152

153

166167

forcedisplacement loops
lead-rubber bearings

176177

triple pendulum bearings

184

velocity-dependent devices

177

178
178179

forcedisplacement spectra
seismic isolation

186

forcevelocity relations

177

foundations

180

2930

capacity design shear


conceptual design

126128
6465

design action effects

206207

248249

detailing

163164

165167

elastic deformations

1314

linear modelling

8690

modal response spectrum analyses

89

90

nonlinear static analyses

116117

seismic design examples

217

218

159164

165167

conceptual design

42

45

elastic response spectra

21

verification

219

free cantilevers

free-standing piers
frequency-dependent terms
frictional devices

21
8889
180185

full dynamic model

67

fully linear behaviour

88

fundamental mode analyses


fundamental mode method (FMM) of analysis

58

9297

196197

67

185

design action effects

248249

design examples

233234

seismic isolation

186191

geometric imperfections

215216

geotechnical aspects
Gerber-type hinges
girders

1936
39
3940

global displacement ductility factors

7374

14

global models
detailing/verification

159

This page has been reformatted by Knovel to provide easier navigation.

8184

155

Index Terms

Links

global response
ductility

1213

gravity

80

8283

119120

Greece

41

42

43

23

105

45

46
grillage models

80

ground
accelerations

1921

displacement

25

flow

166

oscillation

167

33

types

2122

velocity

25

see also soil

H
high design ground acceleration

105

high-damping elastomeric bearings (HDEBs)

176

higher-mode effects

113114

hinges
conceptual design
detailing/verification
hollow circular piers

39
164
55

hollow rectangular piers

5455

horizontal elastic response spectra

2124

horizontal force transfers


horizontal non-seismic actions
horizontal reinforcement of joints

224225
132
8485

horizontally flexible and nearly straight decks

9596

horizontally inflexible and nearly straight decks

9495

hydrodynamic effects

103

48

horizontal seismic components

hydraulic viscous dampers

43

177

180

8486

hysteresis
nonlinear static analyses

114

115

nonlinear time-history analysis

236

237240

seismic isolation

174177

I
immediate use (IU) limit states

immersed piers

8486

impedance

8690

7
115116

This page has been reformatted by Knovel to provide easier navigation.

4950

5556

Index Terms

Links

importance factors
imposed deformations
imposed horizontal loads
in shear dimensioning

8
8283
224
202203

in shear ultimate limit states

220

in-ground hinges

164

in-plane stiffness

80

inclined pile foundations

65

incrementally launched decks

21

individual pier model

9394

inelastic deformations

108

inertia forces

30

inertia loads

96

inflexible and nearly straight decks


inspections

20

9495
49

integral bridges

integral connections
integrity verifications of joints
intermediate design ground acceleration
intermediate movement joints
internal forces

6062
129130
105
109110

138

163

irregular ductile bridges

6768

110111

isolation systems

171192

230

analyses

185191

behaviour families

174185

conceptual design

4445

5253

5556

237240

241

174
design seismic action
examples

174
221226

lateral restoring capability


low-damping elastomeric bearings

191
151155

means

171

nonlinear time-history analysis

236

objectives

171

performance requirements

171173

seismic design examples

232233

verification

243246

IU see immediate use limit states

J
joints
capacity design shear

126128

This page has been reformatted by Knovel to provide easier navigation.

171

Index Terms

Links

joints (Cont.)
clearance lengths

135140

dimensioning

130

horizontal reinforcement

132

integrity verifications

129130

maximum reinforcement

130

minimum reinforcement

130

overlap lengths

135140

relative displacements

109110

seismic design examples

207

stress conditions

129

vertical reinforcement

131

209

221

K
kinematic interaction

88

Krystalopighi bridge (GR)

41

43

laminated elastomeric bearings

47

50

lateral force

96

112113

lateral restoring capability


lateral spreading

81

191
32

33

176177

178

3536

LBDP see lower-bound design properties


LDEBs see low-damping elastomeric bearings
lead-rubber bearings (LRB)
life safety (LS)

67

limit states

57

conceptual design
design examples
detailing

54
219220
121

effective flexural stiffness


modal response spectrum analyses

162163

74
172

shear verifications

247

verification

146

103104

seismic isolation
slope stability

58

173

2930
121

146

162163

see also ultimate limit states


limited ductile behaviour
effective flexural stiffness
seismic design examples
verification

15
100
210221
129

This page has been reformatted by Knovel to provide easier navigation.

164167

Index Terms

Links

linear analyses

67

effective stiffness confirmation


seismic displacement calculations
torsional effects
verification procedures

68

73

107
107110
9899
122123

linear modelling
conceptual design

39

foundations

8690

soil

8690

link slabs

8184

links
conceptual design
detailing/verification
liquefaction

63
140142
3236

adverse effects

3233

assessment

3334

detailing

164167

mitigation

3435

settlements

33

verification

164167

35

liquefied backfill

5960

locked-in bridges

61

70

204

221

longitudinal bars
buckling prevention
longitudinal fundamental mode analyses

196197

longitudinal reinforcement

246247

longitudinal response
conceptual design

4142

loose saturated soils

162

low design ground acceleration

105

low seismicity

16

low-damping bearings
detailing/verification

143

145

149155

low-damping elastomeric bearings (LDEBs)


dimensioning
seismic isolation
lower-bound design properties (LBDP)
detailing/verification
fundamental mode methods
nonlinear time-history analysis

142155
173

175176

232234
144

145

154

237

239

233234
236
241

seismic isolation

173

177

This page has been reformatted by Knovel to provide easier navigation.

241242

Index Terms

Links

LRB see lead-rubber bearings


LS see life safety

M
maintenance
conceptual design
mass

49
72

maximum force estimations

190191

maximum reinforcement in joints

130

medium-dense sands

162

Mesovouni bridge (GR)


minimum longitudinal reinforcement
minimum reinforcement in joints

8486

46
246247
130

MMS see multimode spectrum analyses


modal pushover analyses
modal response spectrum analyses

114
67

displacement calculations

108

nonlinear dynamic analyses

112

seismic design examples

197

modelling

68

7392

68

7392

67118

behaviour factors

6973

effective stiffness

100107

fundamental mode analyses


modal response spectrum analyses
nonlinear analyses
seismic action components
modified historic time-history records
MohrCoulomb failure criteria
momentaxial force interaction diagrams
monolithic connections
conceptual design

9297
67
110117
6869
26
3132
197

199

1112
41

4346

4950

4445

58

5758
detailing/verification

159

effective torsional rigidity

106

movable connections
clearance lengths

135140

conceptual design

43

detailing/verification

156158

overlap lengths

135140

see also elastomeric connections; sliding connections


movement joints

109110

multi-column piers

11

5556

This page has been reformatted by Knovel to provide easier navigation.

56

Index Terms

Links

multi-frame models
multi-sliding surface bearings
multi-span bridges
multimode spectrum analyses (MMS)

39
182185
3839
185

187

N
natural eccentricities
near-collapse (NC) limit states
near-source effects
nearly straight decks
Newmarks equal displacement rule

95
7
25
9496
10

NLTH see nonlinear time-history analysis


nodal inertia loads
non-collapse requirements

96
816

non-concrete decks

150

non-critical components

139

non-encased laminated elastomeric bearings

47

non-laminated elastomeric bearings

47

non-liquefied surface layers

166

non-seismic actions

173

non-slender pier columns


nonlinear analyses

140

224225

58
6768

108117

185186

236242

nonlinear dynamic analyses

6768

108109

111114

nonlinear static analyses

6768

112114

160

185186

236242

nonlinear time-history analysis (NLTH)

132135

O
operational (OP) limit states
out-of-plane flexibility

3940

overlap lengths
movable joints

135140

seismic design examples

208209

overpasses

56

overstrength moments

124126

201

P
p multipliers

117

parallel voids

80

parametric analyses
partial seismic isolation

106
172173

This page has been reformatted by Knovel to provide easier navigation.

160

Index Terms

Links

participation factors

74

7579

peak ground acceleration (PGA)

23

27

182183

184185

517

171173

period shifts

171

172

permanent loads

230

pendulum bearings
performance requirements

PGA see peak ground acceleration


piers
action effects

243246

analyses

7072

capacity design shear


conceptual design

126128
3940

detailing/dimensioning

120

effective flexural stiffness

103

effective torsional rigidity

106

elastic response spectra

21

imposed deformations

225

modal response spectrum analyses

8081

nonlinear static analyses

114

overlap lengths

138

4159

8486

second-order effects

215216

section verifications

246248

seismic design examples

193221

230

42

45

213214

230

246248

stiffness
conceptual design
design examples
of elements

194

seismic isolation

188

piles

910

conceptual design
detailing
modal response spectrum analyses
nonlinear static analyses
verification

64

65

163164

165167

90
114

116117

163164

165167

plastic hinges
analyses

70

design

10

1112

detailing

122128

132134

pile internal forces

163164

seismic design examples

197198

199

200
verification

122128

132134

This page has been reformatted by Knovel to provide easier navigation.

199

200

Index Terms

Links

Poisson assumptions
post-elastic strength degradation
pot-bearings

20
115
47

precast girders

3940

prefabricated girders

8184

prestressed concrete decks


principles

74

7679

prismatic 3D beam/column elements

73

protection, structural

purely cohesive soils

161

purely dry cohesionless soils

161

purely flexural elastic behaviour


pushover analyses

74

80

112114

160

195196

211212

12
6768

py criteria

116117

py curves

167

Q
quasi-permanent traffic loads

119

R
Rayleigh quotients

96

records
artificial time-history

26

27

rectangular
bearings

144148

columns

53

piers

5455

sections

101

ties

134

reduction coefficients

103

167

reduction factors
soil

32

reference return periods

reference seismic action

119

reinforcement
detailing

131

dimensioning

130

section verifications

246248

seismic design examples

203204

relative displacements
reliability differentiation

132

220221

109110
1921

This page has been reformatted by Knovel to provide easier navigation.

230231

Index Terms

Links

replacements
conceptual design
response modification factors
response spectra
response-history analyses

49
10
6768

70

215

216

195

214

6768

retaining walls

33

rheological models

89

rigid deck model

93

rigid foundations

8687

9496

rigidity
concrete decks
rigidly connected abutments
Rion-Antirrion Bridge
roadway joints

106
158159
9

89

207

209

rocking impedance

89

rotational-translational coupling

87

165166

S
safety
life

67

structural

sands

162

saturated soils

161

scattering of waves
scragging

162

28
176

SDoF see single-degree-of-freedom systems


seating

48

secant-to-yield-points
effective flexural stiffness
modal response spectrum analyses

100101
215217

section verifications

246248

seismic actions

89

ductile piers

211

fundamental mode analyses


gravity

103

16

1936

80

second-order effects

elastic response spectra

102

1925
97
119121

seismic isolation

174

spatial variability

2829

time-history representations

2628

see also action effects


This page has been reformatted by Knovel to provide easier navigation.

104

Index Terms

Links

seismic design examples

193249

abutments

217

action effects

197200

217219

236246

248249

axial force

197200

219

246

bearings

207208

221

capacity design

201202

204206

206

206

205

206

206

207

225226

231232

207

207
decks
dimensioning

201203

ductility

193221

effective stiffness

213214

230

flexural verification

205

206

206

foundations

217

218

219

fundamental mode method of analysis

233234

isolation systems

221226

232233

207

209

joints
limit states

219220

limited ductile behaviour

210221

modal response spectrum analyses

221

197

overlap lengths

208209

piers

193221

246248

plastic hinges

197198

199

199

200

230

200
reinforcement
stiffness

203204

220221

194

199201

213214

204206

207

ultimate limit states

219220

verification

197200

self-weight

195

service conditions

173

serviceability

serviceability limit states (SLS)

74

33

35

160163

165

12

46

47

63

settlements
shallow foundations
shear force

126128

shear keys
conceptual design
detailing/verification
modal response spectrum analyses

140
81

shear rigidity

80

shear span

11

71

147148

151153

shear strain

This page has been reformatted by Knovel to provide easier navigation.

72

132133

Index Terms

Links

shear ultimate limit states


shear verifications
shear wave velocity measurements
shrinkage

220
247248
33
196

simulated time-history records


single-column piers
analyses

26

27

1112
72

conceptual design
effective torsional rigidity

5556
106

single-degree-of-freedom (SDoF) systems


analyses methods

67

design

10

elastic response spectra

19

seismic isolation
single-sliding surface bearings
siting of soils

186

9394

95

189190

181182
2930

sizing
elastomeric bearings
pier columns

151155
5759

skewed bridges

138139

skewed decks

9899

slab decks

7374

slack of the link

79

80

4748

5153

63

slenderness
pier columns

57

sliding connections
capacity design effects
conceptual design
dimensioning
seismic isolation
sliding surfaces

128
43
142
180181
181183

slope stability

2930

SLS see serviceability limit states


soils
conceptual design

64

damping

32

elastic response spectra

2122

foundation

2930

linear modelling

8690

liquefaction

33

parameters

166167

3032

pile foundations

164

This page has been reformatted by Knovel to provide easier navigation.

5556

Index Terms

Links

soils (Cont.)
properties

3032

reduction factors
saturated

32
161

seismically active faults


siting of

162

29
2930

spatial variability

28

stability

2930

stiffness

32

strength parameters

3132

solid circular columns

53

solid rectangular columns

53

spatial variability

2829

special isolation bearings

5253

spectral analyses

162

185

spherical sliding surfaces


spirals

181185
133

134

lateral spreading

166

167

nonlinear static analyses

117

springs

SPT see Standard Penetration Tests


square root of the sum of the squares (SRSS) rule

29

92

114

163

stability verifications

64

160162

Standard Penetration Tests (SPT)

22
6768

static analyses
steel decks

9899

111

32

33

34

9297

112117

160

230

44

steel elastoplastic energy dissipators

177

179

stiffness
analyses

100107

conceptual design

42

45

design examples

194

199201

213214

linear analyses

107
174175

188

modal response spectrum analyses


seismic isolation
soil properties
strain hardening

8889
173
32
115

strength
balance with ductility
design

15

nonlinear static analyses


soils

1516
115
3132

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

stress conditions of joints

129

structural protection

structural safety

structural seismic analysis models

229230

substructure verifications

243246

superposition

86

surfacing materials

72

sustainability

symbols

symmetric supports

87

225226

9496

T
target displacement

111

temperature variations

225

tension models

39

terms

109

theoretical eccentricities
thermal actions

95
196

three-span bridges

74

7679

tie-downs

42

45

6768

108109

236242

248

time-history analyses
time-history representations
topographic amplification

227229

196

211212

45

45

2628
25

topping slabs

3940

torsional effects

9899

torsional rigidity

106

total shear strain

147148

traffic loads

185186

119

195

230231
transition periods
transverse reinforcement of piers
transverse responses
travelling wave effect
triple pendulum bearings

10
204
4243
28
182183

truss elements

109

tuned mass dampers

119

Turkey

184185
121

40

twin-blade parts
twisting phenomena

4142

44

54

58

98

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

U
UBDP see upper-bound design properties
UDL see uniform traffic loads
ULS see ultimate limit states
ultimate limit states (ULS)
conceptual design

6
54

design examples

219220

detailing

121123

58
130

146

148149

146

148149

162163
effective flexural stiffness
modal response spectrum analyses

103104
74

seismic isolation

172

shear verifications

247

slope stability
verification

173

2930
121123

130

162163
uncracked flexural stiffness
uniform seismic demands
uniform traffic loads (UDL)
uplifting

100
4143
119

195

48

upper-bound design properties (UBDP)


detailing

232233
144

fundamental mode methods

145

154155
240

235236

nonlinear time-history analysis

236

238

seismic isolation

173

177

verification

144

145

154155

25

33

177180

V
velocity
verification
abutments

155159

bearings

142155

of components

119169

design examples

197200

204206

foundations

159164

165167

isolation system substructure

243246

isolation systems

243246

lateral spreading

164167

limited ductile behaviour

156158

129

linear analyses

122123

liquefaction

164167
This page has been reformatted by Knovel to provide easier navigation.

207

241

Index Terms

Links

verification (Cont.)
nonlinear analyses

132135

seismic links

140142

vertical bars

134

vertical components
elastic spectra
fundamental mode analyses

2425
97

vertical loads

116

vertical reinforcement of joints

131

vertical seismic components

8586

very wide bridge decks

9899

viscous damping

21

volumetric strain assessments

35

Votonosi bridge (GR)

41

22

190191

42

W
wall-like piers

54

walls

33

water

8486

waves

28

Winkler foundations

33

163

Winkler models
lateral spreading
modal response spectrum analyses

166

167

8990

nonlinear static analyses

117

wish-bone shaped elements

109

This page has been reformatted by Knovel to provide easier navigation.

You might also like