You are on page 1of 363

High-Speed Railway Bridges

High-Speed Railway Bridges

Conceptual Design Guide

José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado


Authors All books published by Ernst & Sohn are
carefully produced. Nevertheless, authors,
José Romo editors, and publisher do not warrant the
Alejandro Pérez-Caldentey information contained in these books,
Fhecor Ingenieros Consultores, S.A. including this book, to be free of errors. Readers
Barquillo 23 are advised to keep in mind that statements,
28004 Madrid data, illustrations, procedural details or other
Spain items may inadvertently be inaccurate.

Manuel Cuadrado Library of Congress Card No.: applied for


Fundación Caminos de Hierro
Calle Serrano 160 British Library Cataloguing-in-Publication Data
28002 Madrid A catalogue record for this book is available
Spain from the British Library.

Cover: Riudellots Viaduct, Riudellots de Bibliographic information published by


la Selva (Gerona), Spain the Deutsche Nationalbibliothek
Photographer/Copyright: José Romo The Deutsche Nationalbibliothek lists
Madrid, Spain this publication in the Deutsche
Nationalbibliografie; detailed bibliographic
This book was kindly supported by: data are available on the Internet at
<http://dnb.d-nb.de>.

© 2024 Ernst & Sohn GmbH, Rotherstraße 21,


10245 Berlin, Germany

All rights reserved (including those of


translation into other languages). No part of
this book may be reproduced in any form – by
photoprinting, microfilm, or any other
means – nor transmitted or translated into a
machine language without written permission
from the publishers. Registered names,
trademarks, etc. used in this book, even when
not specifically marked as such, are not to be
considered unprotected by law.

Print ISBN: 978-3-433-03313-5


ePDF ISBN: 978-3-433-61040-4
ePub ISBN: 978-3-433-61039-8
oBook ISBN: 978-3-433-61038-1

Cover Design: Petra Franke/Ernst & Sohn GmbH


using a design by Sonja Frank, Berlin, Germany
Typesetting Straive, Chennai, India

Printed on acid-free paper.


v

Contents

Foreword xv
About the Authors xvii
Acknowledgements xix

1 Introduction to High-Speed Railway Bridges 1


José Romo
1.1 Book’s Content 1
1.2 What is Special About a High-Speed Rail Bridge? 2
1.2.1 Dynamic Amplification and Resonance 2
1.2.2 Rail Traffic Security 3
1.2.3 Passenger’s Comfort 3
1.2.4 Track–Structure Interaction 4
1.3 General Ideas on High-Speed Railway Bridges 4
1.4 Evolution and Trends in High-Speed Bridge Design 6
1.4.1 First High-Speed Bridges 6
1.4.1.1 First-Generation German Bridges 6
1.4.1.2 First-Generation French Bridges 8
1.4.1.3 First-Generation Spanish Bridges 8
1.4.2 Recent High-Speed Bridges 9
1.4.2.1 Recent French Bridges 9
1.4.2.2 Second-Generation German Bridges 9
1.4.2.3 Recent Spanish HSRB 10
1.4.2.4 Bridges for High-Speed Railway Lines in China 10
1.4.2.5 British High-Speed Bridges 12
1.4.2.6 High-Speed Railway Bridges in the USA 12
1.4.3 Conclusions 12
1.4.3.1 Viaducts 13
1.4.3.2 Long-Span Bridges 13
1.5 The Landscape and the Design of High-Speed Railway Bridges 13
1.5.1 The Traveller’s Experience 13
1.5.2 The Bridge in the Landscape 15
1.5.2.1 Long Viaducts with Low Vertical Level 16
1.5.2.2 Long Viaducts with Medium or High Level 16
vi Contents

1.6 Railway Bridges as Landmarks or Icons of a Line 22


1.7 Railway Bridge’s Legacy 23
1.8 Building for the 21st Century 24
1.9 Conclusions 24
References 25

2 Track for High-Speed Bridges 29


Manuel Cuadrado
2.1 Introduction 29
2.2 Specific Criteria for Railway Bridges 29
2.2.1 General Criteria 29
2.2.2 Specific Criteria for High-Speed Bridges 31
2.3 Description of the Track Superstructure 31
2.3.1 Track Components: Definitions, Functions, and Qualities 32
2.3.1.1 Ballast 32
2.3.1.2 Sleepers 32
2.3.1.3 Fasteners 33
2.3.1.4 Rails 33
2.3.1.5 Switches and Crossings 34
2.3.2 Most Important Conceptual Improvements 35
2.3.2.1 Continuous Welded Rail (CWR) 35
2.3.2.2 Track–Infrastructure Interaction: Better Understanding 36
2.3.3 Evolution of the Different Components 36
2.3.3.1 Ballast 36
2.3.3.2 Sleepers 37
2.3.3.3 Fastenings 39
2.3.3.4 Rails 40
2.3.4 Track Options Currently Available for High Speed 41
2.3.4.1 Optimised Ballasted Track 41
2.3.4.2 New Ballastless Track 43
2.4 SLS Related to the Track 44
2.4.1 Dynamic Interaction: Track–Vehicle 44
2.4.2 Track Geometry Quality 46
2.4.3 SLS Verifications Regarding Deformations and Vibrations 48
2.4.3.1 Criteria for Traffic Safety 48
2.4.3.2 Comfort Criteria 55
References 57

3 Conceptual Design of High-Speed Railway Bridges 61


José Romo
3.1 Introduction 61
3.2 Structural and Functional Specific Requirements for High-Speed
Railway Bridges 62
3.2.1 Introduction 62
3.2.2 Control of Vertical Acceleration 62
Contents vii

3.2.3 Rotation at Expansion Joints 62


3.2.4 Horizontal Braking and Traction Forces and Relative Movements
Between Deck and Infrastructure 62
3.2.5 Track-Bridge Deck Interaction 63
3.2.6 Expansion Joints 63
3.3 Longitudinal Design Strategies 64
3.3.1 General Concepts 64
3.3.2 Ballasted Track 65
3.3.3 Ballastless Track 66
3.3.4 Actions to be Considered at the Fixed Point 66
3.4 Design Situation of High-Speed Railway Bridges 66
3.4.1 Short Crossing at Low Level 67
3.4.2 Long Structures 67
3.4.2.1 Low Profile 68
3.4.3 High-Level Viaducts 71
3.4.4 Long Span Structures 72
3.5 Structural Types 72
3.5.1 Straight Deck Solutions 72
3.5.1.1 Simply Supported Deck 72
3.5.1.2 Continuous Slab Concrete Decks 73
3.5.1.3 Precast Beam Decks 74
3.5.1.4 Concrete Box Hollow Sections 78
3.5.1.5 Steel Beam Decks 80
3.5.1.6 Steel Semi-through Decks 81
3.5.2 Truss Bridges 82
3.5.3 Arch Bridges 83
3.5.3.1 Upper Deck Bridges 83
3.5.3.2 Tied Arch Bridges 85
3.5.4 Cable-Supported Bridges 85
3.5.4.1 Extradosed Bridges 85
3.5.4.2 Cable-Stayed Bridges 86
3.5.4.3 Suspension Bridges 88
3.5.4.4 Hybrid Bridges 89
3.6 Structural Elements – Substructure 89
3.6.1 Abutments 90
3.6.1.1 Abutments with Expansion Joint in Structure Only 90
3.6.1.2 Abutments with Expansion Joint in Structure and Track 90
3.6.1.3 Fixed Abutments 91
3.6.2 Piers 95
3.6.3 Bearings 95
3.6.3.1 General Bearing Layout 96
3.7 Seismic Design 99
3.7.1 Seismic Design Strategies 99
3.7.2 Seismic Behaviour and Deck Articulation 99
3.7.3 Longitudinal Behaviour 100
viii Contents

3.7.3.1 Simply Supported Spans 100


3.7.3.2 Continuous Deck 100
3.7.4 Transversal Behaviour 101
3.7.4.1 Introduction 101
3.7.4.2 Fixed Transversal Support 101
3.7.4.3 Transversal Damping Systems 102
3.7.4.4 Damping Devices Plus Bearings 104
3.8 Worked Example 106
3.8.1 Introduction: Aim and Data 106
3.8.1.1 Topography 106
3.8.1.2 Plan and Elevation 106
3.8.1.3 Railway Platform Section – Project Speed 106
3.8.1.4 Water Flood Level 107
3.8.1.5 Preliminary Geotechnical Data 107
3.8.2 Methodology 107
3.8.3 Critical Analysis of Existing Information 107
3.8.4 Determination of the Length of the Viaduct, Selection of the Fixed
Point 108
3.8.5 Span Distribution 109
3.8.6 Deck Pre-dimensioning 109
3.8.7 Pre-design of the Infrastructure 112
3.8.7.1 Fixed Point 112
3.8.7.2 Bearings 112
3.8.7.3 Abutments 112
3.8.7.4 Piers 113
References 115

4 Design Basis 117


José Romo
4.1 Introduction 117
4.2 Design Situations 117
4.3 Rail Traffic Actions and Other Actions Specific of Railway Bridges 118
4.3.1 Permanent Loads 118
4.3.1.1 Self-Weight 118
4.3.1.2 Dead Loads 118
4.3.1.3 Partial Ballast Removal 119
4.3.2 Variable Loads 119
4.3.2.1 Vertical Live Loads 119
4.3.2.2 Traction and Braking Forces 121
4.3.2.3 Centrifugal Forces 122
4.3.2.4 Nosing Forces 123
4.3.2.5 Aerodynamic Actions from Passing Trains 123
4.3.2.6 Thermal Actions 123
4.3.2.7 Bearing Friction 124
4.3.3 Dynamics Effects 124
Contents ix

4.3.3.1 Introduction 124


4.3.3.2 Consideration of Dynamic Effects 125
4.3.4 Railway Vehicle Derailment 125
4.3.4.1 Railway Vehicle Impacts 125
4.4 Application of Traffic Loads on Railway Bridges 126
4.4.1 General 126
4.4.1.1 Load Situations for Structural Design 127
4.4.1.2 Load Situations for Limit State and Associated Acceptance
Criteria 127
4.4.2 Groups of Loads 127
4.4.2.1 Characteristic Values of Multicomponent Action 127
4.5 Traffic Loads for Fatigue 128
4.6 Verifications Regarding Deformation and Vibrations for Railway
Bridges 128
4.7 Worked Example 129
4.7.1 Introduction 129
4.7.1.1 Calculation of Reactions at Bearings: Pre-dimensioning 130
4.7.1.2 Calculation of Forces and Preliminary Design of the Fixed
Abutment 130
4.7.2 Actions 130
4.7.2.1 Vertical Loads 130
4.7.2.2 Horizontal Forces 131
4.7.2.3 Wind Speed 132
4.7.3 Calculation of Reactions at Bearings: Pre-dimensioning 133
4.7.3.1 Vertical Forces 133
4.7.3.2 Centrifugal Forces 134
4.7.3.3 Wind at Unloaded State 135
4.7.3.4 Wind with Live Load 135
4.7.3.5 Reactions in Pier Heads 135
4.7.3.6 Transversal Wind Bearings Reactions 136
4.7.3.7 Loads per Bearings 136
4.7.4 Fixed Abutment Loads 137
4.7.4.1 Introduction 137
4.7.4.2 Loads Transmitted by the Deck 137
4.7.4.3 Forces Acting on the Abutment 138
References 140

5 Dynamic Behaviour of High-Speed Railway Bridges 143


Alejandro Pérez-Caldentey
5.1 Introduction 143
5.1.1 Resonance 143
5.1.2 Envelope Dynamic Factor 144
5.1.3 Dynamic Factor for Real Trains Obtained by Means of Analytical
Formulations 145
5.1.4 Dynamic Factor Obtained by Dynamic Analysis 147
x Contents

5.2 Methods for Dynamic Calculations and Structural Response 153


5.2.1 Modal Superposition 153
5.2.1.1 Matrix Formulation for Finite Element Analysis 153
5.2.1.2 Formulation Based on Assumed Eigenforms 155
5.2.2 Response to the Isolated Load 158
5.2.3 Response to the Train Loads 162
5.2.4 Effect of Damping 164
5.2.5 Dynamic Interaction Between Vehicle and Structure 165
5.3 Interoperability 167
5.3.1 Introduction 167
5.3.2 Universal Dynamic Train A 167
5.3.3 Universal Dynamic Train B 167
5.4 Application Examples 168
5.4.1 Case Without Dynamic Analysis 168
5.4.2 Case with Dynamic Analysis 169
References 183

6 Longitudinal Track–Structure Interaction 185


Manuel Cuadrado and Alejandro Pérez-Caldentey
6.1 Introduction 185
6.2 Problem Statement 185
6.3 Model for Analysis 188
6.3.1 General Considerations 188
6.3.1.1 Rails 189
6.3.1.2 Deck 189
6.3.1.3 Interaction Between Rails and Track Base 189
6.3.1.4 Bearings 189
6.3.1.5 Columns 190
6.3.1.6 Foundations 190
6.4 Actions 191
6.4.1 Temperature Variations 191
6.4.1.1 Case Without Track Joint 191
6.4.1.2 Case with Track Joint 191
6.4.2 Traction and Braking Forces 191
6.4.3 Vertical Loads 192
6.4.4 Creep and Shrinkage 192
6.4.5 Combination of Actions 193
6.5 Verifications 194
6.5.1 Verifications in Terms of Stresses 194
6.5.2 Verifications in Terms of Displacements 195
6.5.3 Criteria for Placing a Track Joint 196
6.6 Rail Expansion Joints 197
6.6.1 Design of REJs – Calculation of the Maximum Displacement 197
6.6.2 Regulation 201
6.6.3 Installation 201
6.7 Longitudinal Schemes 203
Contents xi

6.7.1 Continous Deck with a Single Fixed Point Located at One of the
Abutments 203
6.7.1.1 General 203
6.7.1.2 Examples 204
6.7.2 Continous Deck with the Fixed Point Located on One of the Central
Piers 211
6.7.2.1 General 211
6.7.3 Simply Supported Spans Without Longitudinal Continuity, with a Fixed
Point on Each Span 211
6.7.3.1 General 211
6.7.3.2 Example 212
6.7.4 Fixed Points at the Two Abutments and a Structural Joint in the
Middle 212
6.7.4.1 General 212
6.7.4.2 Example 214
6.7.5 Deck Divided into Several Continuous Stretches, Each One Including
Several Spans and One Fixed Point 216
6.7.5.1 General 216
6.7.5.2 Example 217
6.7.6 Especial Situations 218
6.7.6.1 Seismic Design 218
6.7.6.2 Exceptional Geometries 226
6.7.6.3 Example of Exceptional Geometry 226
6.8 Example of Track–Structure Interaction 229
6.8.1 Verification of Stresses in the Rails 229
6.8.2 Verification of Horizontal Displacement at Abutment 2 Due to Braking
and Traction Forces 231
6.8.3 Verification of Horizontal Displacement at Abutment 2 Due to Vertical
Train Loads 232
6.8.4 Verification of Vertical Displacement at Abutment 2 Due to Vertical
Train Loads and Temperature Variations 234
References 235

7 Conceptual Design for Maintenance 239


José Romo
7.1 Introduction 239
7.2 Accesses 240
7.2.1 Decks 240
7.2.2 Piers 240
7.2.3 Abutments 241
7.3 Bearings 242
7.4 Expansion Joints 243
7.5 Drainage 246
7.6 Conclusions 248
References 248
xii Contents

Appendix A Basic Concepts of Dynamics 249


Alejandro Pérez-Caldentey
A.1 Dynamics of Single Degree-of-Freedom Systems 249
A.1.1 Dynamic Response to Moving Loads (Dynamic Load Factor) 249
A.1.2 Basics of Resonance 257
A.1.3 Solution of the Equation of Motion of a SDOF Damped System
Subjected to a Triangular Load 258
A.1.3.1 Auxiliary Expressions – Integrals I 1 , I 2 , and Their Derivatives 259
A.1.3.2 Solution for the damped SDOF System Subjected to a Triangular
Load 261
Reference 262

Appendix B Singular Bridges for High-Speed Railway Lines 263


José Romo
B.1 Germany 263
B.1.1 Gemünden Bridge 264
B.1.1.1 Data Summary 264
B.1.1.2 Description 264
Further Reading 264
B.1.2 Veitshöchheim Bridge 266
B.1.2.1 Data Summary 266
B.1.2.2 Description 266
Further Reading 266
B.1.3 Pfieffetal Bridge 268
B.1.3.1 Data Summary 268
B.1.3.2 Description 268
Further Reading 268
B.1.4 Nantenbach Bridge 270
B.1.4.1 Data Summary 270
B.1.4.2 Description 270
Further Reading 270
B.1.5 Unstruttal Bridge 272
B.1.5.1 Data Summary 272
B.1.5.2 Description 272
Further Reading 272
B.1.6 Gänsebachtal Viaduct 274
B.1.6.1 Data Summary 274
B.1.6.2 Description 274
Further Reading 274
B.1.7 Hämerten Bridge 276
B.1.7.1 Data Summary 276
B.1.7.2 Description 276
Further Reading 276
B.1.8 Filstal Bridge 278
B.1.8.1 Data Summary 278
Contents xiii

B.1.8.2 Description 278


Further Reading 278
B.2 France 281
B.2.1 Garde-Adhémar Viaduct 282
B.2.1.1 Data Summary 282
B.2.1.2 Description 282
Further Reading 282
B.2.2 Avignon Viaducts 284
B.2.2.1 Data Summary 284
B.2.2.2 Description 284
Further Reading 284
B.2.3 Mornas Viaduct 286
B.2.3.1 Data Summary 286
B.2.3.2 Description 286
Further Reading 286
B.2.4 Savoureuse Viaduct 288
B.2.4.1 Data Summary 288
B.2.4.2 Description 288
Further Reading 288
B.3 Spain 291
B.3.1 Osera Bridge 292
B.3.1.1 Data Summary 292
B.3.1.2 Description 292
Further Reading 292
B.3.2 Llinars Del Vallès Viaduct 294
B.3.2.1 Data Summary 294
B.3.2.2 Description 294
Further Reading 294
B.3.3 Salto Del Carnero Railway Bridge, Saragossa 296
B.3.3.1 Data Summary 296
B.3.3.2 Description 296
Further Reading 296
B.3.4 Viaduct Over AP7 Riudellots de la Selva 298
B.3.4.1 Data Summary 298
B.3.4.2 Description 298
Further Reading 298
B.3.5 Contreras Bridge 300
B.3.5.1 Data Summary 300
B.3.5.2 Description 300
Further Reading 300
B.3.6 Viaduct Over River Ulla 302
B.3.6.1 Data Summary 302
B.3.6.2 Description 302
Further Reading 302
B.3.7 Almonte Bridge 304
xiv Contents

B.3.7.1 Data Summary 304


B.3.7.2 Description 304
Further Reading 304
B.3.8 Alcántara Bridge 306
B.3.8.1 Data Summary 306
B.3.8.2 Description 306
Further Reading 306
B.4 Japan 309
B.4.1 Yashiro Bridge 310
B.4.1.1 Data Summary 310
B.4.1.2 Description 310
Further Reading 310
B.4.2 Kumagawa Bridge 312
B.4.2.1 Data Summary 312
B.4.2.2 Description 312
B.4.3 Sannai-Maruyama Bridge 314
B.4.3.1 Data Summary 314
B.4.3.2 Description 314
Further Reading 314
B.5 China 317
B.5.1 Tianxingzhou Yangtze River Bridge 318
B.5.1.1 Data Summary 318
B.5.1.2 Description 318
Further Reading 318
B.5.2 Nanjing Dashengguan Yangtze River Bridge 320
B.5.2.1 Data Summary 320
B.5.2.2 Description 320
Further Reading 320
B.5.3 Tongling Yangtze River Bridge 322
B.5.3.1 Data Summary 322
B.5.3.2 Description 322
Further Reading 322
B.5.4 Beipanjiang Bridge 324
B.5.4.1 Data Summary 324
B.5.4.2 Description 324
Further Reading 324
B.5.5 Yachihe Bridge 326
B.5.5.1 Data Summary 326
B.5.5.2 Description 326
Further Reading 326
B.5.6 Wufengshan Yangtze River Bridge 328
B.5.6.1 Data Summary 328
B.5.6.2 Description 328
Further Reading 328

Index 331
xv

Foreword

At the request of the authors, I have been given the honour of writing the foreword
to this book, which is devoted to railway bridges. It develops the aspects referring to
their structural conception, taking into account the characteristics of railway traffic:
actions, limit states, speeds, etc., and includes a detailed analysis of the superstruc-
ture of the track with its different components and singular elements (for example,
expansion devices) that allow the correct behaviour of the track.
In the following chapters, the knowledge and experience of the authors is passed
on. In this respect, I remember a technical conference that took place in the 1970s
at the Eduardo Torroja Institute, dedicated to bridges; at that time, the undersigned
engineer was assigned to the Renfe Bridge Division and attended it. Ramón
del Cuvillo, professor of Concrete at the School of Civil Engineering in Madrid,
presented a paper in which he focused on the defects and mistakes in design and
execution in projects and works in which he had been involved. His presentation
was the most applauded of the day’s and, personally, the one from which I learned
the most. I hope that reading this book will be useful to avoid the repetition of
problems that can be avoided, without having to wait for experience after the
execution of the works.
As the reader will appreciate, special emphasis is placed on the interactions
between the structure and the track, subjected to railway and environmental
actions, taking into account the requirements of their stability in different situ-
ations; solutions are also proposed and considered in relation to the transitions
between the bridge and the adjacent infrastructure (and track).
Special attention is paid to the dynamic nature of railway actions, studying the
dynamic response of the structure and its influence on the behaviour, also dynamic,
of the track and its components, with the repercussions that this may have on safety,
traffic flow quality, and maintenance needs.
To conclude, I would like to transmit here some ideas that the Emeritus Professor
of Structural Engineering of the University of Berkeley, Edward L. Wilson, sets out in
his book Static and Dynamic Analysis of Structures. In a section of Personal Remarks,
he relates that his first-year physics professor warned his students ‘not to use an
equation they could not prove’; he also advises, with respect to modern structural
xvi Foreword

engineering, ‘not to use a structural analysis program unless you fully understand
the theory and approximations contained in the program’. I fully agree with these
considerations; I therefore share them with the reader, in the hope that they will be
useful to them.

Madrid, June 2023 Jorge Nasarre


Civil Engineer
Caminos de Hierro Foundation
xvii

About the Authors

José Romo is Chief Executive Officer and partner of FHECOR, and also a bridge
engineer fully specialised in large-span bridges with more than 40 years of expe-
rience in bridge design, 35 of them working in FHECOR. He has vast technical
knowledge based on his design background complemented with his activity as
professor of concrete and steel structures at Madrid University, and his active
participation in national and international associations of bridge designers and
concrete and steel materials. He is a member of many scientific committees such as
Eurocodes, IABSE, and ACHE where he became president in 2014 and was awarded
with the honour’s medal in 2008. He is a fellow of the Institution of Civil Engineers
of UK. He has always worked as a bridge designer participating in innumerable
bridge projects in Spain and worldwide, and also in the construction engineering
for many of them. He has a great aesthetic vision that he applies to all the designs,
while having great concern for sustainability and the use of new materials and
construction techniques.

Alejandro Pérez-Caldentey is full Associate Professor at the Department of


Mechanics of Continuous Media and Theory of Structures for the Civil Engineering
School at the Polytechnic University of Madrid. He joined FHECOR in 1989 after
graduating from UPM where he also obtained his PhD in Civil Engineering in
1996. During his more than 34 years of experience, Alejandro has developed
structural bridge projects in countries such as Spain, Chile, Italy, and the USA. He
is experienced in managing multidisciplinary structural teams, developing designs,
and planning and defining the scope of works. He also has extensive experience in
managing and developing Research and Development projects, in Standardisation
(member of the Project Team for EN 1992-1-1:2023), and in Education (Professor
at UPM). He holds Engineering licenses for Spain, Chile, Virginia, Texas, Florida,
North Carolina, Québec, Ontario, and British Columbia. He is also a partner and
member of the Board of FHECOR Consulting Engineers.
xviii About the Authors

Manuel Cuadrado holds an MSc in Civil engineering from the Polytechnic


University of Madrid. He is currently Associate Professor at the Carlos III University
of Madrid and a member of the Technological Committee of the Spanish Railway
Research Foundation (SRRF). Manuel Cuadrado has been working for 34 years,
mainly in the railway industry, for Spanish and French Engineering companies, as
an independent Consultant, and from 2005 to 2017 for the SRRF. He has partici-
pated both in key Spanish High-Speed projects and in International High-Speed
Lines (Portugal, Turkey, California), and has been involved in many R & D projects
mainly related to the mechanical behaviour of railway infrastructures. As a result
of his R & D activity, he has produced many monographs, published several papers
in national and international journals, and presented many papers in national and
international congresses, including WCRR 1999-Tokyo, WCRR 2001-Köln, WCRR
2006-Montreal, WCRR 2008-Seoul, and WCRR 2016-Milan, and UIC High Speed
Congresses 2010-Beijing and 2015-Tokio. He was also invited to participate as a
specialist in the drafting of railway standards, as a member of Spanish, European,
and international technical committees. Finally, from December 2017, he has been
participating as Infrastructure Assessor and Lead Assessor in several Rail Safety &
Interoperability assessments, as Infrastructure expert and as Slab-track expert.
xix

Acknowledgements

The authors would like to express their most sincere thanks to all the staff of
FHECOR and the Caminos de Hierro Foundation who gave us their support and
assistance in the creation and publication of this book.
We are particularly grateful to Francisco Javier Fernández Pozuelo, for his com-
mitment and dedication to this initiative, and to Eduardo Romo for his support from
the Caminos de Hierro Foundation.
The authors would like to express their deepest gratitude to Jorge Nasarre for his
help in the general approach to the publication and its subsequent technical review,
to Julio Sánchez for his advice and technical reviews, to Fabrice Leray for the prepa-
ration of the graphic material and the conception of the book’s design, to Eduardo
Conde for the layout work, and to Marta Heras for her coordination. We would also
like to thank all the FHECOR engineers who prepared the High-Speed Bridge Design
seminar, which was the seed of this book. Without their help and without the strong
support of FHECOR and the Caminos de Hierro Foundation, this book would not
have been possible. Many thanks to all of them.
1

Introduction to High-Speed Railway Bridges


José Romo

1.1 Book’s Content


One of the particularities of this book is that it includes not only the aspects related
to the design and behaviour of these types of bridges, but also those questions linked
to the railway technology of the track itself. It is clear that the knowledge of both
fields and the interaction between these two technologies, structural and railway, is
fundamental for the complete design of these bridges.
The first chapter of the book is dedicated to explain the particularities of
high-speed railway bridges (HSRB), in comparison with structures for conventional
railways. The typological particularities of this type of bridge are also explained, as
well as the importance of these works as a legacy for future generations.
Chapter 2 is devoted entirely to explaining the technology of the track and the
particularities of the high-speed infrastructure. This chapter explains the special
constraints in terms of rail traffic safety and passenger comfort. It also deals with
critical elements in the design of these structures, such as rail joints and other special
track elements.
Chapter 3 reviews the main concepts which affect the design and includes the
main typologies used in structures for high-speed railway lines. The dimensions and
characteristic weights of the different solutions are also included. This chapter also
describes the special structural elements of these structures, such as abutments and
fixed points. Finally, the particulars of the design of HSRB located in seismic areas
are included. This chapter also has a worked example corresponding to a railway
viaduct, which starts with the general definition of the bridge in a specific valley
and the geometric definition of the different structural elements that make up the
structure.
Chapter 4 is dedicated to the Design Basis of bridges of the railways high-speed
lines. In this section, the typical loads and design criteria are indicated, as well as its
application to the worked example defined in Chapter 3.
Chapter 5 is devoted entirely to analysing the dynamic phenomena associated
with HSR bridges. In this section the different methods of analysis, the trains that

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
2 1 Introduction to High-Speed Railway Bridges

must be analysed to calculate the dynamic response, as well as the way to consider
other aspects of the response, such as the irregularity of the track and the vehicle or
the interaction between the vehicle and the structure, are presented. The chapter
is completed with several practical examples and an appendix which includes
the theoretical aspects of general dynamics and their application to the analysis
of HSRB.
Chapter 6 is dedicated to the interaction between the track and the structure.
This section analyses this phenomenon and how to take into account the thermal
effects, traction and braking forces, vertical loads and rheological effects, in the case
of concrete decks. In addition to the analysis models, the checks to be carried out
to calculate stresses in rails and relative displacements are analysed. This chapter
also deals with the criteria for the placement of track joints, as well as the practical
application of the worked example.
Chapter 7 deals specifically with aspects linked to the conceptual design with
maintenance of bridges for high-speed rail lines in mind.
In addition to Chapters 1–7, the book includes two appendices. One is devoted
to a review of the general concepts of dynamics that the reader of Chapter 5 on the
dynamic behaviour of these bridges should be familiar with. The second appendix
includes a ‘register’ of high-speed railway bridges built in different parts of the world.

1.2 What is Special About a High-Speed Rail Bridge?


It is often asked what is so special about a railway bridge for a high-speed line and
particularly, what makes a railway bridge for a high-speed line different from a con-
ventional railway bridge. The corresponding Sections 1.2.1–1.2.4 that follow in this
chapter describe the causes or aspects that make HSRB so special.

1.2.1 Dynamic Amplification and Resonance


On railway bridges, there are a number of factors that lead to a dynamic response of
the structure under traffic loads.
On the one hand, the loads are fast so there is an impact effect. On the other hand,
the trains are composed of a more or less long succession of vehicles which means
that the loads are repeated, so the dynamic effect is amplified. Finally, the imperfec-
tions of both the track and the vehicles create disturbances in the value and the way
of applying the loads, which leads to an increase in the response of the structure.
Therefore, the actual forces and deformations of a bridge due to rail traffic are
of a dynamic nature and their values can be considerably higher than those due to
static actions. In order to take this amplification into account in the calculations, an
impact or dynamic magnification coefficient is applied to the static loads, a coeffi-
cient established in the design standards on the basis of statistical studies carried out
on bridges in service.
But all these causes are increased when the speed of trains is increased, and as
will be seen throughout the book, the critical range of speeds for the phenomenon
of resonance on a bridge occurs when trains run over 220 km/h.
1.2 What is Special About a High-Speed Rail Bridge? 3

Resonance of a structure occurs when the frequencies of the dynamic excitatory


actions coincide with the eigenfrequency of vibration of the structure f 0 (a whole
fraction of it). In the case of railway bridges, resonance can be produced by the
passage of trains with regularly spaced axle loads or groups of axles (dk metres)
running at a certain critical speed (v in m/s).
v∕dk = f0 ∕i with ; i = 1,2,3 … (1.1)
Thus for a 30 m span bridge with a typical eigenfrequency of 3.5 Hz, on which
high-speed trains with 18 m coaches are running, the critical speed of passage is
3.5 ⋅ 18 ⋅ 3.6 = 227 m/s.
The coefficients of dynamic load magnification do not cover the risk of the effects
of the resonance of the structure.
The amplification of stresses and accelerations due to the proximity to the
resonance frequency means that special problems typical of HSRB can occur.
These problems can affect the functionality of the structure as they can lead on
the one hand to safety problems for rail traffic and on the other hand to a loss of
comfort for train users.
Therefore, it must be verified that the vibrations of the deck do not reduce the
lateral support of the track or reduce the contact pressure between the wheel and
the rail, which could cause the wheel to come off the track and the convoy to derail.

1.2.2 Rail Traffic Security


One of the effects that can jeopardise the safety of rail traffic as a result of the high
speed of the train is the high vertical acceleration of the deck produced as a dynamic
effect of the excitation of the structure if the frequency of the loads is close to the
vertical frequency of the structure. In these cases, track instability can occur as a
result of the loss of ballast support or the loss of geometric quality of the track.
Other effects, such as the danger of derailment by deck twist or by the deformation
of the deck or rotations in supports, or by the transverse deformation of the deck,
or by the relative displacement of the deck, increase considerably as the speed of
passage of train increases.
All this obliges the establishment of much more rigorous limits for the highest
speeds and even, as will be seen later, to create fixed longitudinal connection points
between the deck and the infrastructure to avoid its relative movement.

1.2.3 Passenger’s Comfort


Also, as a consequence of the vertical accelerations suffered by the structure, there
may be a loss of comfort for train users. For this reason, the design of the structure
must seek to distance the vibration frequencies of the structure from the frequency of
passage of the bogies and therefore the loads, in order to reduce this problem so that
the acceleration experienced by the passengers and therefore their loss of comfort is
within manageable limits. To analyse that a dynamic analysis used different types of
trains has to be carried out.
4 1 Introduction to High-Speed Railway Bridges

1.2.4 Track–Structure Interaction


On all railway bridges there is an interaction between the track and the structure.
The track is laid on the structure and therefore there is a joint response to the loads.
For example, the difference in temperature between the rails and the structure, the
transmission of traction and braking loads make it necessary to control the stresses
on the rails to prevent them from breaking. The complexity of the mechanics
of the connection between the rails and the deck and between the deck and the
substructure (including the foundations) means that in any bridge project for a
high-speed line it is necessary to analyse the interaction between the rails and the
structure by means of a non-linear analysis. This type of complex analysis allows
calculating the value of the stresses in the rails as well as the distribution of loads to
the different part of the structure.

1.3 General Ideas on High-Speed Railway Bridges


Here again it might be asked what the differences are between a conventional railway
bridge and a high-speed one. Firstly, it should be noted that the deformation and
acceleration limits that must be met in this type of bridge are much more demanding,
due to the stricter demands on the regularity of the track to achieve high-throughput
speeds, and consequently the decks are slightly more robust than in the case of a
conventional railway bridge.
But perhaps what most differentiates an HSRB from other bridges is the need
to rigidly fix the deck to a fixed point in the infrastructure, in the common case
of continuous decks. This means that on the one hand the longitudinal typology
of these bridges is different and on the other hand the connection details between
superstructure and substructure are special as will be explained below [1]. There is
also another factor that conditions the longitudinal typology. The stroke of the track
expansion devices homologated for high-speed. For a time the maximum stroke
was 600 m and then in the last decades it went up to 1200 mm.
The need to fix the deck longitudinally to one point of the infrastructure, in bridges
with a continuous deck, means that on the one hand the longitudinal behaviour
of this type of bridge is radically different from that of other bridges. Firstly, the
resistance to longitudinal action is concentrated at one point, which means that the
deck will be subject to significant traction and this influences the design of the deck.
On the other hand, when the deck is fixed at one point, it is often necessary to have
rail expansion joints in one of the abutments when the structure exceeds a length of
approximately 90 m in order to reduce the over-stress on the rails.
In all cases where the deck is continuous, at least one element of the infrastructure
must be designed with high longitudinal rigidity (Figure 1.1). As will be seen later,
in the case of long viaducts, and due to the limitation of maximum movements of
commercial expansion joint devices, it may be necessary to have a fixed point in the
middle of the bridge.
An alternative to the previous design is to make isostatic bridges in which each
pier takes the corresponding part of the longitudinal load, especially the traction and
1.3 General Ideas on High-Speed Railway Bridges 5

Figure 1.1 Sar Viaduct (FHECOR), Spain (Source: FHECOR).

Figure 1.2 Span by span isostatic solution: China, China Railways (Courtesy of China
Railways).

braking force. This allows the elimination of joints in the rails but on the contrary,
there are structural expansion joints in all sections of the deck coinciding with the
piles. This solution, which might seem better from the point of view of track mainte-
nance, has the disadvantage of higher maintenance of the structure and in the case
of bridges in seismic areas, the lack of robustness in combination, which could cause
relative movements between adjacent decks during the seismic actions. This type of
design is being used very often in China because it is highly industrialisable and
because it allows for very flexible construction (Figure 1.2).
There are some intermediate alternatives that involve the construction of a series
of continuous sections having intermediate joints every certain number of spans.
This solution has recently been used in Germany [2] and [3], shown in Figure 1.3.
6 1 Introduction to High-Speed Railway Bridges

Figure 1.3 Gänsebachtal Viaduct (schlaich bergermann partner sbp), Germany, DB Netz AG
(Source: Störfix).

In summary, and from a typological point of view, high-speed bridges have some
specific aspects which make them, at least from a longitudinal perspective, different
from other bridges, as will be seen in Sections 1.4 and 1.5.

1.4 Evolution and Trends in High-Speed Bridge Design


1.4.1 First High-Speed Bridges
The first high-speed railway lines were built in the 1960s and 1970s in Japan and
later in Europe in Germany, France, and Spain. The first viaducts for high-speed rail-
way lines were built in Japan. The population density throughout Japan has meant
that most of the lines have been built on viaducts. The typologies of bridges for
the Shinkansen are varied, although the most singular are the extradosed bridges, a
typology that originated in Japan itself.
In Europe, the first high-speed railway lines were built between one and two
decades later than the Japanese lines. On these lines, large sections of embankment
or cuttings alternate with a few viaducts and tunnels at specific points along the
route.

1.4.1.1 First-Generation German Bridges


The German bridges for first-generation high-speed lines were designed according
to the principle of rapid replacement of the decks in case of failure of one of them.
Thus, except in exceptional cases, the bridges were built with isostatic spans, with
maximum spans of 55 m. These bridges were obviously heavier than the continu-
ous bridges and required four bearings and expansion joints in all the pier supports,
which is obviously a problem for the maintenance of the structures. However, this
typology allowed the track to be continuous and therefore without rail joints, with
the great advantage that this entailed for track maintenance.
On these early lines, when the size of the obstacle made it impossible to use iso-
static spans, Deutsche Bahn allowed the use of continuous bridges but with a length
1.4 Evolution and Trends in High-Speed Bridge Design 7

LFC LFC LFC


Horizontal TEJ TEJ Horizontal
fixed point fixed point

Horizontal Horizontal
fixed point fixed point
SEJ 162.5 SEJ 168.0 SEJ 299.0 SEJ 164.0 SEJ
SEJ = Structural expansion joint
TEJ = Track expansion joint
LFC = Longitudinal force coupler

Figure 1.4 Structural scheme bridge over the river Main at Gemünden (1984).

LFC LFC LFC Horizontal


TEJ Horizontal TEJ TEJ
fixed point
fixed point

SEJ SEJ SEJ SEJ SEJ SEJ


237.0 369.5 214.0 160.5 281.6
820.5 442.1
SEJ = Structural expansion joint
TEJ = Track expansion joint
LFC = Longitudinal force coupler

Figure 1.5 Structural scheme bridge over the river Main at Veitshöchheim (1987).

limited to 400 m so that this section could be replaced, at least theoretically, in a


single operation.
The bridges over the river Main in Gemünden (Figure 1.4) and Veitshöchheim
(Figure 1.5) built in 1984 and 1987, respectively, are structures with continuous
spans and therefore with rail expansion joints between the individual deck
subsections [4].
The Pfieffetal Viaduct, built in 1989 (Figure 1.6), was the first bridge in which,
while maintaining the isostatic replaceable span solution, a central point was
designed to collect the braking loads there. The isostatic decks were connected to
each other longitudinally by means of a prestressing system centred on the deck,
which allows the braking loads to be transferred to this central point. In this case,
the track has two expansion joints coinciding with the abutments [4].

Longitudinal force couplers Longitudinal force couplers


TEJ TEJ
Horizontal
fixed point

Structural expansion joints Structural expansion joints


493.0 493.0
TEJ = Track expansion joint

Figure 1.6 Structural scheme bridge over the Pfieffetal Viaduct (1989).
8 1 Introduction to High-Speed Railway Bridges

LFC LFC
Horizontal TEJ TEJ Horizontal
fixed point fixed point

SEJ SEJ SEJ SEJ


447.0 53.0 447.0

SEJ = Structural expansion joint


TEJ = Track expansion joint
LFC = Longitudinal force coupler

Figure 1.7 La Grenette Viaduct with ‘inert’ section and double expansion joints of
structure and track.

Horizontal Horizontal LFC


TEJ fixed point TEJ fixed point TEJ TEJ Horizontal
fixed point

470.0 630.0 456.0


SEJ SEJ SEJ
SEJ
SEJ = Structural expansion joint
TEJ = Track expansion joint
LFC = Longitudinal force coupler

Figure 1.8 Avignon viaducts with intermediate expansion joint (1999).

1.4.1.2 First-Generation French Bridges


In contrast to the first-generation German bridges, the first French high-speed
bridges were built with continuous decks. This implies that for long bridges it is
necessary to provide track expansion joints.
Usually, long decks are divided into three parts. The two lateral sections are long
and are connected longitudinally to their respective abutments. Between these two
spans is an isostatic central span called the neutral or inertial span, which allows
the effective expansion length of the bridge to be divided by two (Figure 1.7). These
bridges have two rail joints coinciding with the two expansion joints of the neutral
portal frame structure [5].
In some cases, the provision of a central expansion joint coinciding with an inter-
mediate point of a span has been tested in order to provide one single track expansion
joint in one intermediate point of and intermediate span instead of the two necessary
in the case of solutions with an inert span (Figure 1.8) [6].
Unlike in other countries, special bridges on French high-speed lines have always
been designed in collaboration with architects and engineers. This has perhaps
meant that all the designs have been special and have somehow departed from
purely structural solutions.

1.4.1.3 First-Generation Spanish Bridges


Spanish bridges built from 1987 are characterised by the use of continuous solutions
with a fixed point, usually at one abutment, and a structural and track expansion
joint at the opposite abutment. In the case of very long bridges, it is common to have
1.4 Evolution and Trends in High-Speed Bridge Design 9

TEJ TEJ
Horizontal
fixed point

SEJ = Structural expansion joint


TEJ = Track expansion joint
SEJ SEJ

Figure 1.9 Example of a Spanish bridge with a continuous deck.

one or two intermediate fixed points in the form of A-shaped piers that transmit the
braking loads to the ground (Figure 1.9).
On the other hand, Spain’s rugged orography has given rise to a series of long-span
bridges in which structural rigour has been combined with the aesthetic and land-
scape aspects of the works, as described below [7].

1.4.2 Recent High-Speed Bridges


1.4.2.1 Recent French Bridges
The bridges on the most modern TGV (Train à Grande Vitesse) lines have generally
maintained the design criterion of long bridges with an inert central span.
A singular bridge in the French high-speed lines is the Savoureuse Viaduct
(Figure 1.10). This viaduct is a succession of short sections with special piers that
have structural expansion joints between them and allow the rail to be continuous
and therefore there is no track expansion joint [8].

1.4.2.2 Second-Generation German Bridges


The bridges already built in Germany in the second decade of the 21st century
have been developed according to the indications of the DB Netz Design Guide for
Railway Bridges in 2008 [2]. With the implementation of this guide, the condition
that the decks should be quickly replaceable has been eliminated and semi-integral
bridges have been designed, generally without track expansion joints and structure
joints every 100–120 m or so. Examples of bridges designed along these lines are the
Unstruttal Bridge and the Gänsebachtal Viaduct.
The Unstruttal Bridge has a height of about 40 m above the valley, the arch spans
have a span of 108 m, and the standard spans have a span of 58 m (Figure 1.11). There
are structural and track expansion joints every 580 m [3].
The Gänsebachtal Viaduct (Figure 1.12) with a deck height of about 18 m, has
shorter spans (24.50 m) and structural expansion joints every 112 m and does not
need track expansion joints [3].

Structural expansion joints


42.0 24.0

66.0

Figure 1.10 La Savoureuse Viaduct (2011).


10 1 Introduction to High-Speed Railway Bridges

TEJ TEJ TEJ TEJ TEJ


HFP HFP HFP HFP HFP HFP
580.0
SEJ SEJ SEJ SEJ SEJ

SEJ = Structural expansion joint


TEJ = Track expansion joint
HFP = Horizontal fixed point

Figure 1.11 Unstruttal Bridge (2012).

Horizontal fixed points

112.0
Structural expansion joints

Figure 1.12 Gänsebachtal Viaduct (2012).

1.4.2.3 Recent Spanish HSRB


On the other hand, Spain’s rugged orography has given rise to a series of long-span
bridges in which structural rigour has been combined with the aesthetic and
landscape aspects of the works, as described in [7] including:
1. Precast decks with constant depth with spans up to 45 m
2. Precast decks with variable depth with spans up to 60 m
3. Prestressed concrete box-section bridges built with movable scaffolding system
(MSS) with spans of up to 75 m
4. Steel–concrete composite deck up to 66 m of span
5. Extradosed steel bridges with spans up to 75 m
6. Bow-string arches with spans in the range of 125–130 m
7. Concrete box deck built by cantilevering up to 150 m of span
8. Steel truss decks with 240 m of span
9. Concrete arches with a maximum span of 384 m
For the above-mentioned bridges see Appendix B.

1.4.2.4 Bridges for High-Speed Railway Lines in China


The structures that have been and are currently being built in China are grouped into
two characteristic types. On the one hand, there are the long multi-span viaducts and
on the other hand, special bridges with one or more main spans with long spans.
The long viaducts currently being built in China are isostatic spans that are prefab-
ricated completely and assembled directly from previously built spans (Figure 1.13).
On the other hand, the orography of mountainous areas and the width of riverbeds
have led to the construction of long-span bridges. The construction of arch bridges,
cable-stayed bridges, and even the first suspension bridge are undoubtedly a sign
of China’s determination to build a high-speed network that does not stop at any
technical challenge.
For bridges mentioned above see also Appendix B and Table 1.1.
1.4 Evolution and Trends in High-Speed Bridge Design 11

Figure 1.13 Isostatic multi span bridge, China (Courtesy of China Railways).

Table 1.1 List of the major high-speed railway bridges (HSRB).

Design
Name of the Main Year of Operation
Bridge Span (m) Typology Completion (km/h) Function* Country

Wufengshan 1092 Suspension 2020 250 4R+8H China


Yangtze River
Hutong Yangtze 1092 Cable stayed 2020 250 4R+6H China
River
Tongling Road-Rail 630 Cable stayed 2015 250 4R+6H China
Anqing Yangtze 580 Cable stayed 2015 250 4R China
River
Huanggang 567 Cable stayed 2014 250 2R+4H China
Yangtze River
Tianxingzhou 504 Cable stayed 2008 250 4R+6H China
Zhaoqing Xi River 450 Steel box arch 2014 250 2R China
Beipan River 445 Concrete arch 2016 300 2R China
Wuhu Yangtze 588 Cable stayed 2019 250 4R+6H China
River
Yachi River 436 Steel truss arch 2019 250 2R China
New Baishatuo 432 Cable stayed 2019 250 6R China
Yangtze River
Almonte River 384 Concrete arch 2016 330 2R Spain
Dashengguan 336 Truss arch 2011 300 6R China
Yangtze River
Yibin Jinsha River 336 Bow-string arch 2018 250 4R+6H China
Alcántara 324 Concrete arch 2019 330 2R Spain
Grümpen 270 Concrete arch 2011 300 2R Germany
Contreras 261 Concrete arch 2009 350 2R Spain
Dongping Channel 242 Truss arch 2009 NA NA China
Ulla Estuary 240 Truss beam 2015 250 2R Spain
*
R = number of railway tracks.
H = number of highway lines.
12 1 Introduction to High-Speed Railway Bridges

Figure 1.14 Colne Valley Viaduct of HS2, England, UK (Courtesy of Knight Architects).

Figure 1.15 Fresno Viaduct in California, USA (Source: California High-Speed Rail
Authority).

1.4.2.5 British High-Speed Bridges


The HS2 line between London and Birmingham is currently under construction.
This line runs on flat terrain and the viaducts to be built have a very low gradient. The
solutions being used in general are prefabricated multi-girder decks with continuity
lintels constructed on site.
The Colne Valley Viaduct (Figure 1.14), which crosses one of the most scenic
areas of the line, stands out as a singular bridge within the section. It is a contin-
uous viaduct with maximum spans of 105 m that has a very careful design and will
undoubtedly be the most emblematic bridge on the line.

1.4.2.6 High-Speed Railway Bridges in the USA


The high-speed rail line from San Francisco and Sacramento to Los Angeles is
currently under construction. The structures being built are generally continuous
viaducts with concrete box section or prefabricated girders. One of the most
important works on the line is the Fresno River Viaduct (Figure 1.15).
The other line in the project phase is the line from Dallas to Houston, which is
also planned with solutions using concrete box-section decks and precast concrete
elements.

1.4.3 Conclusions
In order to establish an analysis of trends in the design and construction of
high-speed bridges, it is first necessary to make a clear distinction between viaducts
1.5 The Landscape and the Design of High-Speed Railway Bridges 13

of more or less length on the one hand, and bridges requiring a long main span on
the other hand.

1.4.3.1 Viaducts
There are currently three types of solutions for viaducts with moderate spans and
shorter or shorter lengths.
In Europe, with the exception of Germany, continuous bridges are generally built
with a minimum number of structural expansion joints and corresponding track
expansion joints. The deck is supported by two devices per pier and per abutment.
The bridges on the US lines also generally follow these design criteria.
In China, on the other hand, isostatic solutions are being built with a complete
prefabrication in one piece of the deck of each span and therefore with an expansion
joint on each pier, but without track expansion joints.
As seen in Germany, semi-integral bridges are being designed and built, i.e. with-
out pier bearings, but with structural expansion joints, and in some cases without
track expansion joints.

1.4.3.2 Long-Span Bridges


The typologies being used to date are similar to road bridges. Deck trusses with
variable edge have been used up to 250 m span. For longer spans (up to 450 m), arch
bridges have been successfully built. For longer spans, cable-stayed bridges are the
usual solution. In these cases, the decks are trussed to give greater rigidity to the
system and often have two levels: the lower one for rail traffic and the upper one
for road traffic.
Table 1.1 summarises the bridges for high-speed lines with the longest spans built
to date.

1.5 The Landscape and the Design of High-Speed


Railway Bridges
1.5.1 The Traveller’s Experience
The 21st century society, at least in the West, is governed by feelings [9]. The quality
of any service is measured by the user experience [10]. High-speed rail is no excep-
tion to this premise.
Traditionally, when the designer analysed the engineering work and its relation-
ship with the landscape, they did so taking into account only the view of the observer
of the bridge from the surrounding environment. In this way, it is common to anal-
yse from the different points from which the bridge or viaduct can be observed what
modification it will introduce into the pre-existing landscape. The height and config-
uration of the abutments, the cadence of the deck spans, the relative dimensions of
the deck and its spatial relationship with the morphology of the site are the aspects
on which the bridge designer concerned with the landscape reflects.
14 1 Introduction to High-Speed Railway Bridges

1 m high parapet for


maintenance

Noise barrier panels


with partial or complete
opacity to avoid fow
and bat impact
where needed

OLE posts on the external


side of the deck following
the modulation of the piers

What a HS2 user would see from the window of the train if a solid noise barrier is
arranged (static image on the left and train running at a speed of 340 km/h on the right).
Users wouldn’t even notice they are crossing the Colne Valley

Figure 1.16 Example of a standard anti-noise panel on the bridge (Courtesy of Knight
Architects).

However, it is clear that this is no longer enough. The 21st century bridge designer
must also consider the landscape that the traveller will be able to contemplate when
the train travels over the bridge being designed [11].
When this aspect is analysed, it is discovered that the structure itself rarely
obstructs the view from the train in any way. However, it is common that parts of
the bridge equipment, especially the anti-noise or wind barriers (when these are
opaque), disturb or limit the view of the landscape from the train (Figure 1.16).
If the bridge is short, the loss of vision caused by such panels would only be for
a few seconds. However, when the tracks run continuously through urban or
peri-urban areas, the tunnel effect can be annoying or uncomfortable for the user.
The same applies when the railway line passes through a point of outstanding
scenic beauty, such as a major river crossing, if the passenger’s view of the outside
is limited by some element of the bridge.
In such cases, it will be important for the viaduct designer to be aware of whether
the structure requires any type of panelling that will at least partially obstruct
the vision of the traveller. Whether it is necessary to install panels or whether
it is the structure itself that is disturbing, for example if the resistant section of
1.5 The Landscape and the Design of High-Speed Railway Bridges 15

3 m high linear elements act as a fowl


and bat protection barrier and as a
support for transparent panels when
a noise barrier is needed

1 m high linear elements act as a


parapet for maintenance

Transparent noise barrier


panels where needed

OLE posts on the external


side of the deck following the
modulation of the piers
Concrete surface with vertical
grooves to guide water coming from
the noise barrier/parapet/fowl and
bat protection system

What a HS2 user would see from the window of the train
with the transparent Specimen Design barrier (static
image on the left and train running at a speed of 340 km/h
on the right). At train speed, motion blur makes the vertical
elements of the edge condition almost completely invisible,
achieving an unobstructed view of the Colne Valley.

Figure 1.17 Study of the view from the train as it passes over the Colne Valley Viaduct,
England, UK (Courtesy of Knight Architects).

the deck is U-shaped, the project team must analyse whether it is possible to
reconcile functional requirements (noise emission control, wind safety, etc.) with
the possibility of the traveller being able to enjoy the landscape at least for a fleeting
glimpse (Figure 1.17).

1.5.2 The Bridge in the Landscape


The railway layouts of the 19th century and those built later for moderate traffic
speeds allowed for the adaptation of the railway line to the orography, except in
mountainous areas. However, compared to roads, railways have always needed more
bridges and viaducts to overcome the natural obstacles they have encountered, as
the layout conditions have been and still are more rigorous in the case of railways
compared to the design requirements of roads.
The railway has transformed and continues to transform the landscape through
which it passes. High-speed lines with their very wide radii of curvature in plan
of about 8000 m minimum for 350 km/h lines require the construction of a large
16 1 Introduction to High-Speed Railway Bridges

number of viaducts and tunnels as soon as the terrain has some movement. Even
in flat terrain, it is common for modern high-speed lines to be built in structure in
order to maintain transverse territorial permeability under bridges. It is very impor-
tant in these cases to decide correctly on the level of the railway grade on the ground
because of its implications for the design of the viaducts.
The participation of bridge specialists in the early stages of the project is important
for the definition of the basic geometry of the line and to avoid starting the project
with initial conditioning factors that could damage the overall quality of the solution,
for example: the transverse permeability, the landscape implications of the design,
the technical quality, or the construction cost of the work.
Sections 1.5.2.1 and 1.5.2.2 analyse the landscape aspects of bridges and viaducts
on high-speed lines in different scenarios.

1.5.2.1 Long Viaducts with Low Vertical Level


If the level is relatively low in relation to the ground (less than 8 m), the spans have to
be short in order to leave a sufficient clearance between the bottom of the deck and
the natural ground. If it is also necessary to install noise barriers, the height of the
noise barrier must be added to the actual height of the deck under the track, which
will make the bridge visually very heavy. One way to solve this problem is to use ‘U’
sections, with the structure itself acting as a noise barrier so that the clear span is as
large as possible (Figure 1.18).
However, whenever possible, it is better to raise the level somewhat to avoid
the aforementioned problems. From the point of view of the cost of the bridge, an
increase in the height of the piers from 8 to 12 m has little influence on the final
cost of the structure, since there will only be a slight variation in the foundations
and a higher cost of the pier shafts (which in any case is a small cost in relation
to the total cost of the bridge) and this increase in height will not condition the
construction process of the deck. In any case, and whenever it is necessary to build
either noise barriers or U-shaped structural sections, it is essential to study the
shapes and finishes in order to break up the massiveness of the faces.

1.5.2.2 Long Viaducts with Medium or High Level


As explained above, when viaducts are long, it is necessary to fix the deck to the
infrastructure. If the bridge is also very high, the connecting element(s) will play a
special role in the formal appreciation of the bridge in the landscape.

Figure 1.18 Semi-through deck structure (Source: FHECOR).


1.5 The Landscape and the Design of High-Speed Railway Bridges 17

Figure 1.19 Isostatic deck, China (Courtesy of China Railways).

Figure 1.20 La Savoureuse Viaduct (2011), France (Courtesy of Wilkinson Eyre).

Isostatic Bridges When the viaduct is made up of a succession of isostatic spans, it is


common for both the deck itself and the piers to be particularly robust. As the decks
are isostatic, they are less efficient than continuous decks and require a greater depth
(see Chapter 3). The piers also have to individually withstand the corresponding
braking load and therefore require larger dimensions than in the case of continuous
structures (Figure 1.19).
The La Savoureuse Viaduct (Figure 1.20) has recently been built, breaking with
the French tradition of continuous bridges. In this viaduct, the piers are formed by
a tetrapod-shaped structure supported at one of its vertices, which on the one hand
breaks the massiveness of the piers of isostatic bridges and on the other hand reduces
the span of the isostatic spans. The result is a unique structure that works well in the
surrounding views.

Continuous Bridges Continuous bridges have the advantage of reducing the number
of expansion joints in the structure. When these bridges are long, they require one
or more points to fix the deck longitudinally.
The first example of this way of solving the central connection by means of a sin-
gle element is the Pfieffetal Viaduct in Germany, 1989 (Figure 1.21). This bridge is
actually an isostatic span bridge, but because of its height the piers cannot carry the
braking load, which is transferred to a portal pier with two inclined piers. The shape
of the V-shaped valley makes the role of this central pier very clear.
18 1 Introduction to High-Speed Railway Bridges

Figure 1.21 Pfieffetal Viaduct (1989), Germany (Courtesy of Wolfgang Pehlemann).

Figure 1.22 Bridge over the river Main at Gemünden (1984), Germany (Courtesy of
Deutsche Bahn AG).

Another type of situation occurs when a long viaduct is required which can be
resolved with modest spans, but which presents a singular span due to having
to cross a major obstacle locally. This type of solution perhaps begins with the
Gemünden Bridge (Figure 1.22) , which serves as the cover of the most widely read
book on bridge aesthetics [12].
However, to return to very long viaducts that require a single span, the revolution
brought about in Germany by the Deutsche Bahn Guide [2] is worth mentioning. It
stipulates that long bridges for the Deutsche Bahn should generally be semi-integral
and as far as possible without a track expansion joint.
The first long bridges designed according to these guidelines are the Unstruttal
(Figure 1.23) and Gänsebachtal (Figure 1.24) viaducts [3]. Both bridges are superb in
terms of design, structural efficiency, maintenance of both bridge and track, as well
as structural innovation, with an obvious reading on the landscape to the trained eye.

Singular Bridges Another classic design situation occurs when the obstacle to be
overcome is significant and it is necessary to build at least one large span. This is
a situation that occurs when crossing deep valleys or when passing over very wide
and fast-flowing rivers or streams.
When crossing deep valleys, it is common for the bridge to be a short interval
between tunnels. This is the case, for example, with the colossal Beipanjiang Bridge
(Figure 1.25) on the high-speed line from Shanghai to Kunming in the Chinese
1.5 The Landscape and the Design of High-Speed Railway Bridges 19

Figure 1.23 Unstruttal Bridge (2012), Germany (Courtesy of Deutsche Bahn AG).

Figure 1.24 Gänsebachtal Viaduct (2012), Germany (Courtesy of Störfix).

Figure 1.25 Beipanjiang Viaduct (2016), China (Courtesy of China Railways).


20 1 Introduction to High-Speed Railway Bridges

Figure 1.26 Alcántara Bridge (2019), Spain (Courtesy of Carlos Fernández Casado CFC &
ADIF).

Figure 1.27 Almonte Viaduct (2016), Spain (Courtesy of Arenas Asociados & ADIF).

province of Guizhou, which with its 445 m main span fits perfectly into the nar-
rowing of the gorge flanked by two tunnels [13].
In areas with a hilly but not necessarily mountainous topography, it may be neces-
sary to build a large main span accompanied by two long access viaducts. This is the
case, for example, of two Spanish bridges, the Alcántara viaduct (Figure 1.26) with a
span of 324 m [14] and the Almonte viaduct (Figure 1.27) with 384 m main span [15].
In both cases, the continuity of the deck and the careful design not only of the
main span but also of the access viaducts stand out, making these works excellent
examples of the concern for the visual aspect typical of Spanish engineering.
The above three are good examples of how well rigid arches fit into a rugged land-
scape when there is a major obstacle.
It should be pointed out here that, although at first sight the arch is a typology
which is not very suitable for use on the railway as it has to support significant vari-
able loads, in long-span bridges, the weight of the structure itself is the dominant
1.5 The Landscape and the Design of High-Speed Railway Bridges 21

gravity load compared to the railway overload and it is sufficient for the arch to be
sufficiently rigid for it to function correctly.
Another special design situation for high-speed line bridges is the crossing of
large rivers or estuaries. Especially when the crossing depths or navigation demands
require long spans.
The solutions in these cases are usually trusses, either with straight girders (for
moderate spans) or bow-string or cable-stayed bridges for large spans.
The Nantenbach Bridge (Figure 1.28) in Germany [16] and the Ulla Bridge
(Figure 1.29) in Spain [17] are good examples of the application of the variable-edge
straight-deck typology at river or estuary crossings. In both cases they are bridges
for two high-speed railways tracks.
When it is a question of covering very long spans and with many railway tracks,
the solution is usually a double-level cable-stayed bridge deck. This is the case, for
example, with the Hutong Bridge over the Yangtze River (Figure 1.30). The bridge

Figure 1.28 Nantenbach Bridge (1993), Germany (Courtesy of Deutsche Bahn AG).

Figure 1.29 Ulla Estuary Viaduct (2015), Spain (Courtesy of ADIF).


22 1 Introduction to High-Speed Railway Bridges

Figure 1.30 Hutong Yangtze River Bridge, China (Courtesy of China Railways).

Figure 1.31 Proposal Terceira Travessia do Tejo, Lisbon, Portugal


(Source: ADF-FHECOR-IDEAM).

has a colossal main span (1092 m) in which the upper part of the deck serves a motor-
way while the lower part accommodates the high-speed railway.
In both cases the piers are bottle-shaped. An alternative to these piers is the one
proposed for the Third Tagus River Crossing in Lisbon (Figure 1.31), also designed
with a two-level deck. In this case, the pylon has A-shaped outer shafts, which give
the pylon a very clean form despite its robustness.

1.6 Railway Bridges as Landmarks or Icons of a Line


Fortunately, the requirements for bridges on high-speed lines are so stringent that
only relatively short-span bridges have been built to date, and only very rarely, with
designs that go against the logic of strength or function. Formal speculation is there-
fore far removed from long spans but sometimes manifests itself in long viaducts
with short spans where the structural challenge is not so relevant.
Undoubtedly, at least in Europe, there is a particular attention to aesthetics.
Aesthetics not as a branch of philosophy that studies the essence of beauty and the
perception of beauty in art, but aesthetics as the image of any object.
This ‘aesthetisation’ of the world, as Lipovetsky reminds us [18], strives to make
the formal expression of any object novel or unique. Bridges are no strangers to this
drive, and their design has been influenced in some way, in these first decades of
the 21st century, by this trend in which the most important thing is the appearance
of form.
1.7 Railway Bridge’s Legacy 23

Fortunately, in the case of long-span or long bridges, the site and the functional
requirements they have to meet make each bridge a unique and unrepeatable
example. This should at least calm the desire for novelty [19].
This does not mean that designers should abandon the care for the formal aspects
of their bridges. On the contrary, they should strive to combine appearance and sub-
stance. An effort that should be directed not only towards achieving excellence in
singular works, but also towards seeking beauty, following Yanagi’s ideas [20], in
normal bridges, those that are seen every day.
The expert eye and perhaps also the non-specialist will appreciate the effort made
in many high-speed bridges to wisely reconcile these in no way contradictory aspects
of design. These exemplary works will surely become part of the collective heritage
and culture of the technical community and society in general.

1.7 Railway Bridge’s Legacy

Since its beginnings, the railway has marked the evolution of bridges. The condi-
tions required by the 19th century railway marked the development of structural
engineering, both in terms of materials and in the typology of the bridges themselves
(Figure 1.32).
It is the epic era of engineering marked by successes but also by great failures. The
dynamic loads due to the railway, the fatigue of the materials gave rise to resounding
accidents that served as a lesson to move towards safer structures.
The magnitude of the challenges and the integrity of the engineers had a strong
impact on the society of that time. The material and human development that
was achieved thanks to this new means of transport revolutionised the society of
that time.
Society as a whole and art did nothing but focus on the railway and its works.
Bridges appear in paintings first and then in photography creating a whole imaginary
regarding the progress created by the engineers.

Figure 1.32 Maria Pia Bridge (Ponte de Dona Maria Pia, 1878) Porto, Portugal
(Source: Seyring [21], photo: José Olgon).
24 1 Introduction to High-Speed Railway Bridges

Figure 1.33 Proposal for the Ulla Bridge, Spain (Source: FHECOR).

Perhaps the railway bridges where the first works in which society could admire
the plastic and formal value of the engineers’ work. The visual strength of these
structures undoubtedly had a favourable impact on a society that admired the value
of the technique with fascination [22].

1.8 Building for the 21st Century


High-speed rail routes have very high radii of curvature in plan. This requires the
construction of long tunnels and viaducts when the terrain is hilly. Unlike railways
laid out in the 19th and 20th centuries, bridges and viaducts are no longer the main
defining elements of the route. The bridges are subordinated to the route and this
means that they are very long constructions which undoubtedly shape a modelled
landscape.
These bridges are therefore an excellent opportunity to showcase the advances in
structural engineering of the 21st century. In addition to the technical challenges
of this type of construction, there are also the demands of a mature society that is
logically concerned about the environment and the sustainability of any project.
Structural engineering has to rise to these great challenges. In this way, both the
design of the bridges, and therefore their formal expression, and their construction
and setting in the terrain must be respectful of their landscape and environment [23].
These designs, especially when it comes to singular works, must be a sample of
engineering concerned with the maintenance and durability of their structures,
creating bridges that can be the legacy of the engineers of the first decades of the
21st century (Figure 1.33).

1.9 Conclusions
HSRB are the best example of how a technology, the railway, is developed to reach
land speeds only previously imagined by man. The speed and the necessary safety of
the lines require highly demanding structures that have to meet strict design criteria.
The structural engineering of the moment has very powerful design and calculation
tools that allow the design of these bridges that were unthinkable a few decades ago.
References 25

However, in order to continue achieving challenges, collaboration between railway


engineers and bridge designers is essential. The following text is a good example of
this joint knowledge that is offered to the reader here.

References

1 Manterola J., Astiz M.A., and Martínez A. (1999). Puentes de Ferrocarril de Alta
Velocidad Revista de Obras Públicas n∘ 3386 Madrid.
2 Schlaich, J., Schmitt, V., Marx, S. et al. (2008). Leitfaden Gestalten von
Eisenbahnbrücken, 1e. Berlin: DB Netz AG.
3 Schlaich M. (2012). Integral Railway Bridges in Germany. 22nd Dresdener
Brückenbausymposium. Dresden.
4 Zellner W. and Saul R. (1991). Long span bridges of the new railroads lines in
Germany. IABSE Report 64.
5 Ramondenc P. and Bousquet C. (2004). The main bridges of the high speed line
HSL Méditerranée FIB Sympsium Avignon.
6 Chatelard P., Martin O., Roujon M. and Sayn P. (1998). Lot 2H – Les viaducs
d’Avignon. In: Travaux, n. 742.
7 Recuero, A. (2014). Viaductos singulares del Siglo XXI – Ferrocarril. Revista del
Ministerio de Fomento.
8 Yazbeck, N. (2010). LGV Rhin-Rhône, Branche Est – Les études du viaduc de la
Savoureuse. In: Travaux, n. 870.
9 Russell K. (1983). The Age of Sentiments Modern Age; Chicago, Ill. Vol. 27,
Iss. 3.
10 Dewey, J. (1934). Art as Experience. New York: Minton, Balch & Company.
11 Knight, M.; Beade, H. Colne Valley viaduct specimen design. Government UK
Publications.
12 Leonhardt, F. (1983). Bridges Aesthetic and Design. Deutsche Verlags-Anstalt
GmbH.
13 Chen, B., Su, J.Z., Lin, S. et al. (2017). Development and application of concrete
arch bridges in China. Journal of Asian Concrete Federation 3 (1): 12–19.
14 Manterola, J., Martínez, A., and Martín, B. (2015). Viaducto sobre el río Tajo
en el embalse de Alcántara para ferrocarril de alta velocidad. Revista de Obras
públicas n∘ 3562 Madrid.
15 Arenas, J.J., Capellán, G., Martínez, J. et al. (2016). Viaduct over River Almonte.
Design and analysis. Presented at:. In: IABSE Symposium: Challenges in Design
and Construction of an Innovative and Sustainable Built Environment, Stockholm,
Sweden.
16 Zellner W. and Saul R. (1991). Railway bridge with double composite action
across river Main IABSE Report 64.
17 Millanes, F., Ortega, M., and Matute, L. (2014). Viaduct over river Ulla: an
outstanding composite (steel and concrete) high-speed railway viaduct. Structural
Engineering International 24 (1): 131–136.
26 1 Introduction to High-Speed Railway Bridges

18 Serroy, J. and Lipovetsky, G. (2013). L’esthétisation du monde: L’esthétisation du


monde. Éditions Gallimard.
19 Qin, S. and Gao, Z. (2017). Developments and prospects of long-span high-speed
railway bride technologies. Engineering 3 (6): 787–794.
20 Yanagi, S. (2018). The Beauty of Everyday Things. London: Penguin Classics Art
& Design.
21 Moreira M., Lopes J.M., and Gomes R.M. et al. (2005). Ponte Maria Pia.
A obra-prima de Seyring. Ordem dos Engenheiros Região Norte. Matosinhos.
22 Gandil J. (1991). Evolution des ouvrages des lignes TGV en fonction des
contraintes. Annales ITBTP n∘ 497.
23 Romo, J. (2015). Conceptual Design and Aesthetic of Bridges. In: IABSE
Conference Elegance in Structures IABSE Report V.104 Nara, Japan.
29

Track for High-Speed Bridges


Manuel Cuadrado

2.1 Introduction
Track presents a high level of interaction with the infrastructure components. This
deep relationship means that the optimum design is reached considering the civil
work components, as earthwork or bridges, and the track together along the design
process.
Both the short-term response of the infrastructure under traffic loads and
its long-term behaviour have an impact on railway ride safety and ride quality,
especially in the case of high-speed lines.
In this sense, it must be pointed out the importance of controlling the vertical track
stiffness, related to the short-term combined response of track and infrastructure,
and the evolution of track geometry quality, related to the long-term behaviour and
finally their influence on vehicle dynamic and, ultimately, on ride safety and ride
quality.
In the case of bridges, the design criteria related to the track have a high impact
on the final design of the structure. The present chapter is focused in describing the
track for a better understanding of this design impact.

2.2 Specific Criteria for Railway Bridges

2.2.1 General Criteria


Related to railway bridges, the first question that we can ask ourselves is: What is
the main difference between railway and other similar structures, as road bridges?
Loads are usually mentioned as one of the most relevant differences, and it is
indeed true. In the case of the magnitude of vertical loads, dead loads, due to the
presence of track superstructure, are multiplied by four compared to road bridges
pavement (Figure 2.1).
Also traffic loads, due to the high weight of trains, are multiplied by three com-
pared to road bridges traffic (Figure 2.2).

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
30 2 Track for High-Speed Bridges

Dead loads: x4

Figure 2.1 Differences in dead loads between road bridges and railway bridges.

Traffic loads: x3

Figure 2.2 Differences in traffic loads between road bridges and railway bridges.

That justifies the lower slenderness depth/span of railway bridge decks (about 1/12
to 1/15, depending on longitudinal scheme and the construction method, instead of
1/20 or lower of road bridges).
But not only are vertical loads higher, but also horizontal, due to transverse cen-
trifugal and nosing forces and longitudinal traction and braking forces are higher.
Beyond the magnitude of loads, it is also important the dynamic character of loads,
since it can produce resonance in the case of railway bridges, as it will be presented
in Chapter 5.
On the other hand, the fact that traffic loads are relatively important compared to
permanent loads makes the bridge subject to many cycles of high stress variability,
which can lead to fatigue problems.
But the question of what the main difference between railway bridges and road
bridges is would be still open, because the abovementioned difference in loads is
not the only significant difference. In fact, it might be said that it is not the most
important one.
The design of a railway bridge must never forget the importance of the interactions
between the bridge, rail, and train to result in an adequately functioning system. Due
to that interaction, vibrations shall be limited to ensure both track stability and ride
quality and safety; movements shall be limited to guarantee again a proper level of
track stability and track geometry quality, as well as a limitation of efforts on the
different track components; these aspects are critical.
Thus, the set of verifications related to the track, regarding deformations and vibra-
tions for railways bridges are extremely important, a set of limitations in which
criteria for traffic safety and ride quality are clearly identified. These criteria will
be presented in Section 0.
2.3 Description of the Track Superstructure 31

Figure 2.3 SLS for the bridge – ULS for


track and vehicle.

It must be pointed out that these limitations related to the track, regarding defor-
mation and vibration, are Serviceability Limit State (SLS) for the bridge, but Ultimate
Limit State (ULS) for the track and vehicle, as they are related to running safety
(Figure 2.3).

2.2.2 Specific Criteria for High-Speed Bridges


As described in Section 2.2.1, traffic loads are especially constraining in the case of
railways bridges. One of its main important characteristics is its dynamic nature,
that can eventually produce resonance.
The risk of resonance, even though not exclusive to the high speed, can be
considered as especially vital in the design of high-speed bridges so much so that
international standards, as Eurocode EN 1991-2 [1], distinguishes between bridges
for speeds higher or lower than 200 km/h in the conditions related to the need of
dynamic analysis, as it will be seen in Chapter 5.
However, this aspect cannot be considered as the most important in the specific
design of high-speed bridges, but again the conditions coming from the interaction
with the track. Limitations of vibrations and, especially, of track geometry defect
induced by the deformations of the bridge, can become especially critical in the case
of highspeed.

2.3 Description of the Track Superstructure

Conventional track superstructure consists in ballast, sleepers, fasteners, rails,


switches, and crossings and other complementary equipment.
Operational conditions have experienced major changes during the last decades.
It could be highlighted the increase on the speed for passenger services and the heav-
ier axle loads for freight trains. These facts, together with more dense traffic, have
required the technical evolution of the track components to get higher levels of geo-
metrical accuracy and reliability.
Because of that, the different components mentioned before have evolved refin-
ing their design, developing the selection and control process of the used elements,
and finally, applying new materials which have been appearing along time when
technology has enabled to produce them with more competitive prices.
A review of the evolution of different track’s components is included below.
32 2 Track for High-Speed Bridges

2.3.1 Track Components: Definitions, Functions, and Qualities


2.3.1.1 Ballast
Ballast is defined as the aggregate formed by stones or crushed and screened rock,
according to pre-established requirements.
It has as main functions:

– softening the actions that vehicles cause on the track and transmit them to the
subgrade (distribute uniformly the load and vibrations, softening them, longitu-
dinally and transversally, from the sleepers to the subgrade);
– limiting the movements of the track, stabilising it on the vertical, longitudinal,
and transverse direction;
– providing elasticity and capacity to absorb energy to the track, making it a softener
layer for higher ride quality;
– enabling drainage of rainwater;
– protecting the subgrade materials from the action of the freeze;
– avoiding leakage of electric current (track insulation depends on ballast bed con-
tamination);
– enabling the establishment and recovery of track geometry through levelling,
aligning, and achieving the cant track, by tamping operations;
– reducing the noise generated by the running of the trains.

The following qualities must be reached:

– elasticity enough to absorb the actions of vehicles and to distribute the loads;
effective resistance to prevent horizontal displacements of the track;
– appropriate porosity rate (without affecting elasticity) to allow storm water
evacuations;
– stable properties to deal with the action of water and ice;
– be easily compacted by mechanical means to support elastically the loads trans-
mitted by the sleepers and allow recovering the track initial geometry.

2.3.1.2 Sleepers
They are the cross-structural elements that, resting on the ballast, provide support
to the rails.
Sleepers have as fundamental missions:

– establishing and maintaining the track gauge;


– maintaining the track stability by the appropriate absorption and transmission of
traffic loads across its supporting surface;
– providing the rails slope (1/20) in their supporting area, whereby improving their
stability and the rotating and guiding of the wheels;
– keeping restraint the rails, on height (lifts and settlements) as well as against lat-
eral and longitudinal efforts (braking, acceleration, and thermal efforts);
2.3 Description of the Track Superstructure 33

– additionally, they contribute to preserve the road running surface during construc-
tion, in case of break of the rail, after derailments, and to absorb vibrations of the
track and lower the acoustic impact on the environment.

2.3.1.3 Fasteners
It defines a piece or a group of pieces of a track system that fix the rail to the sleepers
and keeps them in the required position, while preventing (or partially allowing)
vertical, cross, and longitudinal movements.
The performances required to these elements are of three types: (i) mechanical,
(ii) electrical, and (iii) functional.

Mechanical Allocate, absorb, and/or transmit the traffic loads (static and dynamic)
to the sleeper and keep rail attached to the sleeper with enough tightness in all pos-
sible cases.
Thus, these mechanical functions are:
– avoiding the overturning of rails (torsion);
– setting the rails against their support to restrict lateral movements and inclination
keeping the track gauge within a permissible tolerance;
– contributing to the vertical elasticity of the track together with the ballast, function
fundamentally allocated to the clips and elastic pads;
– decreasing the dynamic effects from the rails to the sleepers through the impact’s
absorption in the vertical direction of the loads and vibrations induced by the pas-
sage of trains, mission assigned to the elastic pads;
– preventing longitudinal sliding of the rail on the sleeper;
– controlling the forces on the rails caused by temperature changes (contraction and
expansion).

Electrical Providing adequate electrical insulation (electrified lines or with electrical


signalling).

Functional
– preventing the components wear on the contact surface with the sleeper;
– getting a natural frequency of vibration greater than the rail one, to prevent
contact lost between rail and wheel;
– elastic deformation enough to maintain an elastic reaction under maximum defor-
mation.

2.3.1.4 Rails
The rail is the longitudinal metallic element (steel), which physically constitutes the
road running surface of the train wheels. Trains run along on these elements with
steel wheels, using the key feature of the railway that is the low friction between both
components.
34 2 Track for High-Speed Bridges

Rail must have several features that allow it to withstand a complex set of efforts,
which must fit its profile, its length, and its metallurgical composition.
It is therefore essential to ensure its optimum quality and conservation, which
involves a systematic and continuous analysis of its flaws, cracks, and wears.
The quality requirements are higher when the rail lines support passenger traf-
fic, particularly in the case of high-speed services. They are linked to the following
functions:
– unidirectionally guiding the rolling material with maximum continuity, both in
plan and elevation;
– serving as absorber, resistant and transmission element of the thermal stresses
and traffic loads – vertical, transverse, and longitudinal – to the sleeper and, thus,
to the infrastructure;
– acting as a conductive element for returning the electric current of the vehicle’s
traction, as well as the electric supply of signalling circuits;
– not rapidly aging by weather (oxidation).
To withstand stresses of different nature and high value, rails must achieve the
following requirements:
– high wear resistance;
– high compressive strength;
– high fatigue resistance;
– high yield, tensile strength, and hardness;
– high tear resistance;
– good welding properties;
– high degree of purity of the materials;
– good quality of the rolling contact surface;
– good flatness and fidelity profile;
– reduced internal strains after the manufacturing process.

2.3.1.5 Switches and Crossings


They are the appropriate devices to ensure track continuity when connecting differ-
ent lines, under safe running conditions.
These elements are:
– point switch: ensures the continuity of each of the divergent itinerary;
– common crossing assembly (acute crossing): allows the intersection of two rails
from opposite sides;
– double crossing (obtuse crossing): allows the intersection of two rails from the
same side.
The following type of devices can be distinguished:
(a) Turnouts: it is a device that allows to rail traffic the diversion of a track into two
or three tracks, without interrupting the progress, whose axles are tangential to
the first (or at a very small angle with it).
(b) Crossover: it is a track configuration based on two turnouts, establishing com-
munication between two parallel or concentric tracks.
(c) Diamond crossing: it is a device that allows the intersection of two tracks.
2.3 Description of the Track Superstructure 35

(d) Rail expansion joints: is a track device allowing the relative longitudinal
movement between two adjacent rails while ensuring correct guidance and
support. These longitudinal movements can appear: at the end of continuous
welded rail (CWR), at the end of the structures, or at a combination of the two
precedent situations.

2.3.2 Most Important Conceptual Improvements


Regarding the historical development of the track, it is worth to point out some
concepts that have helped significantly in the great evolution of the system in recent
decades.

2.3.2.1 Continuous Welded Rail (CWR)


Until the appearance of this system, in which the track is made continuous by weld-
ing of the rails, track consisted in rails with a given length joined to one another by
joints with flanges or other connection elements as seen in Figure 2.4.
Along the passage of the wheels, this kind of joints tends to bend as a cantilever
beam, causing an impact that:
– increases tensile strength of trains;
– allows the horizontal shift of the rails;
– produces crushing of the ballast;
– imposes deformations on the rail that may become permanent;
– produces lack of ride quality due to the discontinuity on the running surface;
– produces fatigue in the rolling stock.
For these reasons all operations related to the conservation of their joints were
multiplied:
– the levelling by tamping the ballast;
– the lubrication and tightening of flange bolts;
– checking and controlling the opening of the gaps between rails;
– the consolidation and adjustment of the rail to the sleeper;
– the maintenance of the sleepers;
– the placement and adjustment of rail anchors, etc.

Figure 2.4 Joints of the rails (source [2]).


36 2 Track for High-Speed Bridges

Such problems result in poor geometry of the track and rapid wear of the elements
which form the joints, reducing the speed of circulations and causing a lack of ride
quality, which would be lower, the smaller the lengths of the rails between joints are.
Considering these problems caused by rail joints, in the early twentieth century
CWR began to be used. It was first introduced in trams, which had rails embedded
in the pavement and thus hindered their movements on rails and tunnels which
suffered minor variations in temperature. Today their use has been fully generalised
with the improvement of design criteria and welding techniques.

2.3.2.2 Track–Infrastructure Interaction: Better Understanding


Another important development has been the consideration and analysis of the inte-
grated infrastructure–track system. In this sense, the main conclusion is that their
different properties, in terms of resistance and mechanical deformation, thermal
deformation, and properties related to durability and fatigue, must be analysed.
This has given rise to the study of track–bridge interaction, considering the track
and the deck as well as the stresses that are generated by deformations of different
nature. Longitudinal track-bridge interaction will be analysed in Chapter 6.
Another concept is the influence of the global vertical track stiffness, and espe-
cially the influence of the variation of this stiffness in the deterioration of the system
and hence its impact on the maintenance needs.

2.3.3 Evolution of the Different Components


2.3.3.1 Ballast
The ballast is the only element of the track that has not been modified over time.
It is the product of the crushing of natural rock with appropriate characteristics
(Figure 2.5).
Its evolution has occurred around the following four aspects:
– the specifications that are required with respect to their mechanical properties
have been set and refined;

(a) (b)

Figure 2.5 Ballast (a) and ballast detail (b) (source [2]).
2.3 Description of the Track Superstructure 37

– very demanding constraints on granulometric composition have been defined to


improve its performance and durability;
– design methods and calculation of ballast bedding depending on the type of lines
used have been enhanced. Thus, the minimum thickness required under sleepers
and other dimensions (as the ballast shoulder) have been refined and enhanced;
– the maintenance criteria and the study of lifetime have also been refined and
systematised.

2.3.3.2 Sleepers
Sleepers, being an artificial element designed specifically for railway function,
have evolved in parallel with the advance in technology and advances in the
understanding of the behaviour of the railway infrastructure.
The main types of sleepers that have developed throughout history are presented
below. It must be noted that all these types are still used in different worldwide
networks.

(a) Timber sleeper: Are wood elements, prism-shaped, with highly variable length
depending on the gauge and approximate cross-sectional setting section 25 cm
wide by 14 cm high. Those elements are protected against external actions by
impregnation with a product derived from the distillation of coal oil (Figure 2.6).
(b) Steel sleepers: Generally inverted U-shaped. These sleepers are light,
around 80 kg of weight, and have deterioration problems by oxidation in wet
environments. They allow an industrial design which drastically increases the
homogeneity between the different units and at the same time allows designing
rail support areas so that they are fixed (Figure 2.7).
(c) Bi-block concrete sleepers: With the emergence of quality-controlled and
suitable for prefabrication concrete bi-block prefabricated sleepers were
developed. They have two blocks of reinforced concrete and a steel stay that
maintains a fixed separation between the blocks. They have little horizontal
and vertical rigidity which reduces the rigidity of the track. Also, they bring the
problem of corrosion (Figure 2.8).
(d) Monobloc concrete sleepers: Finally, this is the most used typology in mod-
ern railways. They are made of reinforced concrete and prestressed concrete. Its
main advantages are the bigger weight, which increases the lateral stability, and
higher transverse and vertical rigidity and durability (Figure 2.9).

Figure 2.6 Timber sleepers (source [2]).


38 2 Track for High-Speed Bridges

Figure 2.7 Steel sleepers (source [2]).

Figure 2.8 Bi-block sleepers (source [2]).

Figure 2.9 Monoblock sleepers (source [2]).

(e) Special sleepers: There are innovative designs that have been developed for
special situations or have not been successful because they produce significant
changes in the processes of construction, maintenance, and renewal of rail
infrastructure. One of the most important examples would be frame sleepers
developed in Russia. These sleepers achieve increased vertical and transversal
stiffness, thereby increasing the lateral stability. Besides the longitudinal
stiffness is homogenised because rail is supported longitudinally (Figure 2.10).
2.3 Description of the Track Superstructure 39

Figure 2.10 Frame sleepers (source [2]).

2.3.3.3 Fastenings
There are several types and classifications based on different criteria.
One of the most important classifications is direct, in which the rail and the
sleeper are directly connected through rail-pads and indirect, in which between
the rail and the sleeper there is the baseplate system, which includes a rail-pad and
a baseplate-pad.
Secondly, fasteners can be classified into rigid, semi-elastic, and elastic.
The most used types of fasteners based on both classifications are the following:
(a) direct rigid (with nail or lag-screw);
(b) direct elastic (Figure 2.11);
(c) indirect elastic (Figure 2.12).

(a) (b)

Figure 2.11 Elastic direct fastening. (a) Nabla, (b) Pandrol Fastclip (source [2]).
40 2 Track for High-Speed Bridges

(a) (b)

Figure 2.12 Elastic indirect fastening. (a) Pandrol. (b) Vossloh (source [2]).

2.3.3.4 Rails
The rail is made of steel since XIX. Initially it was given a profile with variable depth
gauge between supports made of stone dices (the ‘fish belly’ rail), leading to the first
type of rail that deserves to be mentioned (Figure 2.13).
The manufacturing difficulty of these rail resulted since then to prism-shaped
profiles to roughly double T. Numerous profiles that were discarded rapidly until
the appearance of double-head rail Bull-Head whose section had double symmetry
were developed. Thus, manufacturing is simplified and it could be reused turning
it. However, when trying to reuse, the lower head was warped which prevented it
reuse. Further advances in scientific knowledge about rolling forced a joint design
with the profile rail in coordination with the profile of the wheels resulted in
the appearance of the Vignoles-type rails, composed of three distinct parts, head,
web, and base. Some types of Vignoles rail type standardised by UIC are shown in
Figure 2.14.

Figure 2.13 ‘Fish bally’ rail.


2.3 Description of the Track Superstructure 41

(a) (b)

Figure 2.14 Vignoles rails; (a) UIC 46 E1; (b) UIC 60 E1.

Figure 2.15 Grooved rail.


Railhead

Groove
Checkrail

Web

Foot

Finally, for cases in which compatibility of rail and road traffic must be assured,
appropriate rails, as grooved rails, to be embedded in the general pavement are used
(Figure 2.15).

2.3.4 Track Options Currently Available for High Speed


2.3.4.1 Optimised Ballasted Track
The ballasted track has evolved over nearly 200 years. Throughout all this time the
different components of the track have been improved as above mentioned.
This process, actually, has led to modern high-performance lines to use a type of
railway superstructure virtually identical in all countries that have this type of rail-
way lines. This type of superstructure over all these years has confirmed an excellent
level of adaptation to the required functions, as well as a long-term performance and
42 2 Track for High-Speed Bridges

Figure 2.16 Ballasted high-speed track. (a) Track in execution. (b) Madrid – Valladolid line
(source [2]).

4.650
4.000
3.350
1.818
1.100

E.E VIA
0.06

3
0.45 0.40 0.80 2
5% Ballast
0.30 Sub-ballast
5%
0.29

2
1 0.60
Blanket
5%

Figure 2.17 Cross section of a ballasted high-speed track (source [2]).

a reliability perfectly adapted to the life of railway lines which is estimated in about
100 years (Figures 2.16 and 2.17).
It also allows the maintenance and renewal processes, processes that can be per-
formed over the lifetime of the line without major affection to operation of the lines.
A typical ballasted track section in the case of a modern high-speed line would
consist of:
– heavy Rail, the UIC 60 type (60 kg/m) that allows a good distribution of verti-
cal and lateral forces. The geometric quality of the rails and their welds must be
especially careful to avoid the effects of wave formation;
– prestressed monoblock concrete sleepers with a length ≥ 2.50 m to ensure good
anchorage of the track in the ballast;
– the fastener would be formed by anchors, elastic pads, and resilient clips type
Skl-1 or similar, that guarantee, without maintenance, an excellent conservation
of tightening torque and an adequate vertical stiffness;
2.3 Description of the Track Superstructure 43

– a ballast with high hardness and important thickness (30 cm beneath sleeper),
which ensures a good resistance to wear and crushing, produced as a result of
dynamic overloads.

2.3.4.2 New Ballastless Track


Ballastless track can be defined as the configuration in which the ballast has been
replaced by a concrete slab, asphalt base, or a metal base.
With this type of track the available experience is much more limited. The ini-
tial experiences of unballasted track applied to modern lines date from the 1970s
in Japan and has not been until the last 15 years when has been started to be used
significantly in high performance networks (Table 2.1).
It is important to highlight that in the Japanese case the characteristics of rail traf-
fic on high-speed lines are very different from the rest of the world. The number of
transported passengers and the very high frequency of circulation necessarily have
led the slab track being generally used. To ensure the reliability of the operation, with
frequencies almost typical of subways and high speeds, and with the very limited
time available for maintenance it was decided long before for this type of solution.
On the other hand, the economic criteria that are used in lines carrying such a large
demand are not applied in other countries.
Apart from the case of Japan, in the rest of the world this technology is beginning
to develop and has led to many different systems trying to lower the cost and simplify
its installation. The main countries that have started using this type of superstructure
are Japan, Germany, China, Taiwan, and Holland.
Functional performance of ballastless track is excellent. It supports higher loads
and once the rail is fixed no loss of levelling or alignment is expected. It is therefore
predicted major maintenance is no longer needed to maintain the track geometry
quality.

Table 2.1 First executions of ballastless track.

1924 Japan early experiences


1966 Japan prefabricated concrete slab
1972 Germany Rheda station: sleepers joined by
longitudinal reinforcement anchored to a
concrete slab on two rigid subbases
1970s France and systems Stedef and PACT, respectively
Great Britain
1980s Theoretical and experimental development
of slab track. Implementation in local trains,
bridges, tunnels, underground, stations.
1990s Development of systems for high speed.
Germany decided to build HS with slab track.
44 2 Track for High-Speed Bridges

Rail on Rail on
resilient baseplate resilient baseplate
Derailment Derailment
upstand upstand
(optional) (optional)

Reinforced concret
e track slab
Bridge deck /
tunnel invert sla
b / etc.

Figure 2.18 Typical slab track on bridges (source [2]).

The concrete slab may be built ‘in situ’, or using precasted parts. The asphalt base
is constructed under continuous compaction.
In some cases not only the ballast has been replaced, but also the sleepers. All
functions of the ballast and/or sleepers must be assumed by some of the components
of the ballastless track. The aim is to replace ballast by other material assuring a
greater stability track.
Thus, the fundamental components of the ballastlees track are:
– rail in long welded bars with elastic fastener (special for ballastless track);
– main slab (reinforced concrete or asphalt, on which rests the track formed by
sleepers, fasteners, and rails);
– base slab (can be gravel-cement or mass or reinforced concrete);
– subgrade (including protective layer and eventually anti-frost layer);
– anti-noise elements (optional).
In the case of ballastless track on a bridge, the deck will substitute the base slab
and the subgrade, as shown in Figure 2.18.
The classification of the different types of existing ballastless track can be made
according to Figure 2.19.
This classification considers on the one hand the elastic characteristics of the sys-
tem, and on the other the features of rail support.

2.4 SLS Related to the Track


2.4.1 Dynamic Interaction: Track–Vehicle
As already pointed out in Section 2.2, the design of a railway bridge must consider
the dynamic interaction between the bridge and the vehicle, to result in an ade-
quately functioning system. In Section 2.4.2 it will be seen in detail how to ensure
2.4 SLS Related to the Track 45

Families
SLABTRACK CLASSIFICATIóN
1 2 3 4 5
Other Fixation Continuous Discrete
characteristics system
Rail levelling Every rail independently Using precasted elements
Precast
No No or block Sleper Slab
elements
Constructive Embedded Sleepers on Precast
Embedded rails Without sleepers
Elastic levels method sleepers support layer slab
EDILON PACT
1 elastic level on CDM porotrac TRANOSA JNR*
fastening TRANOSA
THYSSEN KRUPP
THYSSEN KRUPP
A Züblin BTE RHEDA
Sistema IVESA Heilit W. BES RHEDA berlin ATD BÖGL
2 elastic levels on APPITRACK RHEDA 2000 BTD ÖBB-
fastening CrailsheimFCC RHEDA Diwidag GETRAC PORR*
Rasengleis HEITKAMP IPA*
Hochtief/SM ZÜBLIN
NS blokkenspoor
Elastic level SONNEVILLE STEDEF
between block or B Bloques EDILON SATEBA
sleeper and slab Bloques TRANOSA
RAIL TECH

Figure 2.19 Classification of existing ballastless track (source [2]).

both track stability and ride quality and safety by the limitation of deformations
and vibrations of the track.
Regarding railway vehicle dynamics and then ride safety and passenger comfort, it
is interesting to review the classic Prud’homme formula for computation of dynamic
vertical overloads for short wavelength track defect:

𝜎ΔQ = V ⋅ A ⋅ mun ⋅ k
where
𝜎 ΔQ is the standard deviation of dynamic overload;
V is the running speed;
mun is the unsprung mass (wheelset);
k is the track vertical stiffness;
A is a parameter dependent on track geometry quality.
Thus, these vertical overloads are dependent on the speed, the track geome-
try quality, the unsprung mass, which is the mass that vibrates with the track
(the wheelset, considering that the rest of the mass of the train is connected to the
wheelsets through suspensions that represent a dynamic isolation), and finally the
vertical track stiffness. Track stiffness and track geometry quality will be then two
fundamental parameters of the track–infrastructure integrated behaviour.
One of the main objectives of track structural design reviewed in Section 2.3
is to provide adequate vertical stiffness of the track. A certain value of vertical
flexibility is required to distribute loads, and because one of the main aspects of
the mechanical behaviour of the track system is its dynamic response, and so, the
46 2 Track for High-Speed Bridges

aim is to minimise the level of vibrations in order to reduce dynamic overloads,


so to achieve a lower rate of deterioration and to assure running safety and ride
comfort. Finally, to insulate vibrations from the environment if required. But, on
the other hand, the flexibility is to be limited for practical reasons, to limit stress
in certain track components (mainly rails and fastenings), to assure track stability,
because too much vertical displacements might favour irreversible deformation,
to favour riding stability and comfort and to prevent rail tilting and track gauge
variations.
To establish a parameter that is independent of the individual components, the
vertical stiffness is defined as the ratio between the wheel force and the total vertical
displacement of top of rail in the point of action of the wheel force (Figure 2.20).
The common value for vertical displacement of rail is in the range of 1–2 mm,
and so state of the art is that the vertical track stiffness must be in the range of
50–100 kN/mm. The total stiffness depends on the flexibility of all the flexible com-
ponents: rail pad, sleeper pad, ballast, and subgrade. Related to that it must be con-
sidered that there are two main available track systems for high speed, as seen in
Section 2.3, the classical ballasted track and new solutions without the ballast as a
component that have appeared as new technical option.
Section 5.2.5 reviewed how to consider this parameter for the analysis of the
dynamic interaction between vehicle and structure.

2.4.2 Track Geometry Quality


The second main parameter that influences the appearance of dynamic overloads
and can reduce traffic safety and comfort is the track geometry quality. The
importance of knowing this parameter arose in the middle of 20th century, when
European infrastructure managers developed their own track recording vehicles
allowing a continuous measurement of track geometry and, based on this, their
own track geometry quality evaluation standards.

Hypothetical
Q wheel force = 100 kN

Rail

Rail pad
Z
Sleeper CT describes
Sleeper pad the stiffness below
the wheel relative
CT Ballast or concrete slab
to a wheel force
Ballast mat
of 100 kN
Track stiffness
Substructure/subgrade or
concrete slab/base

Figure 2.20 Vertical track stiffness (source [3]).


2.4 SLS Related to the Track 47

For track geometry quality in Europe the specification in use is the standard EN
13848 [4, 5]. The main purpose of the standard is to define a minimum track geom-
etry quality to ensure safe operation of trains based on the experience of various
infrastructure managers.
Three indicators can describe the track geometric quality:

– extreme values of isolated defects, considered as mean to peak values;


– standard deviation over a defined length, typically 200 m;
– mean value.

Four types of irregularities are considered:

1. Alignment, that is the average of the lateral positions of two rails.


2. The cross-level, that is the difference between the elevations of two rails.
3. Track gauge, defined as the horizontal distance between the two rails.
4. Longitudinal level, that is the average elevation of the two rails.

Alignment, cross-level, and track gauge variations are the major causes of lateral
vibration in railway vehicles, whereas vertical profile is the more important parame-
ter for vertical vibrations, through the parameter A in the Prud’homme formula seen
in Section 2.4.1.
Three main levels must be considered to make decision about maintenance
actions:

● Immediate action limit (IAL): refers to the value which, if exceeded, requires
taking measures to reduce the risk of derailment to an acceptable level.
● Intervention limit (IL): refers to the value which, if exceeded, requires correc-
tive maintenance.
● Alert limit (AL): refers to the value which, if exceeded, requires that the track
geometry condition is analysed and considered in the regularly planned mainte-
nance operations.

And, on the other hand, it is intuitive that not only the value of the defect is impor-
tant, but also the length in which we have the value. That it is not the same, a defect
of 1 mm in one meter than in one thousand meters. To describe a geometric error by
its wavelength (𝜆), three ranges are used:

– D1: 3 m < 𝜆 < 25 m


– D2: 25 m < 𝜆 < 70 m
– D3: 70 m < 𝜆 < 150 m.

Thus, limit values are given as a function of the wavelength range considered and
of speed, being the latter an important factor for the evaluation of track geometry
quality. As an example, we can see in this table the alert, intervention, and immediate
action limits for longitudinal level isolated defects (Table 2.2).
It can be seen how low those limit values are. For instance, for high speed a defect
of only 6 mm in 25 m is already an alert limit.
48 2 Track for High-Speed Bridges

Table 2.2 Limit values for longitudinal level – Isolated defects (Mean to peak value)
(source [5]).

AL (mm) IL (mm) IAL (mm)


Wavelength range Wavelength range Wavelength range

Speed (km/h) D1 D2 D1 D2 D1 D2

V ≤ 80 12 to 18 N/A 17 to 21 N/A 28 N/A


80 < V ≤ 120 10 to 16 N/A 13 to 19 N/A 26 N/A
120 < V ≤ 160 8 to 15 N/A 10 to 17 N/A 23 N/A
160 < V ≤ 220 7 to 12 14 to 20 9 to 14 18 to 23 20 33
220 < V ≤ 300 6 to 10 12 to 18 8 to 12 16 to 20 16 28

2.4.3 SLS Verifications Regarding Deformations and Vibrations


As abovementioned, the deformations, displacements, and accelerations in railway
bridge that affect the track installed in them are SLS for the bridge, but ULS for the
track or for vehicles that circulate, as safety is affected.
In standards, as Eurocode EN 1990 Annex A2, those SLS are included and the
values that must be respected not to affect safety are defined.
Engineers dedicated to the inception, execution, and operation of railway bridges
should not forget that the bridge must support another structure, the track (with
or without ballast), to which the actions applied to the bridge are transmitted,
and whose correct conditions and safe functionality must be checked in all
circumstances.

2.4.3.1 Criteria for Traffic Safety


Vertical Acceleration of the Deck After commissioning the first high-speed line
in Europe (Paris–Lyon line) it was detected the appearance, in some bridges, of
disorders of the ballast bed that caused the destabilisation of the track and the
deterioration of track geometry quality, with the consequent risk for traffic.
A thorough study of this phenomenon revealed that it was associated with vertical
accelerations of the deck, caused by the passage of trains at certain speeds (resonant
speeds that, consequently, they did not have to be the maximum one), when reaching
values of the order of 0.7–0.8 g. The bridges had been designed according to the UIC
71 load model and respecting the rules of UIC 776-1R [6].
In addition, when cracking appears on concrete bridges, the rigidity and frequency
of the deck decreased and the corresponding resonance velocity too. For instance, a
decrease in resonant speed has been detected from 287 km/h (for a new bridge) to
250 km/h for the same bridge after having been put into service.
Once analysed this issue at theoretical and experimental levels, it has been found
that:
(a) In the ballast, subjected to vertical acceleration of the order from 0.7 to 0.8 g, a
phenomenon similar to liquefaction appears that produces the loss of its bearing
capacity and causes a degradation of the levelling and alignment of the track.
2.4 SLS Related to the Track 49

(b) The placement of an elastic mat between deck and ballast amplifies the acceler-
ations in the ballast and therefore is harmful for this phenomenon. This impor-
tant conclusion has been obtained from the results of the tests commissioned by
the ERRI D214 committee [7].
(c) Accelerations of the order of g, even on decks without ballast, can decrease the
contact forces between rail and wheel (even causing take-offs) to inacceptable
limits.
(d) Regular and repetitive distribution of axes of the branches high speed can
result (at certain speeds) to resonant situations of the decks with important
amplifications both in vertical deformation and accelerations, as can be seen in
Figure 2.21.
Consequently, to ensure traffic safety, where a dynamic analysis is necessary
(usually for high speed), the verification of maximum peak deck acceleration due to
rail traffic actions shall be done. The goal is to avoid ballast instability on ballasted
bridges and unacceptable reduction in wheel rail contact forces on slab track bridges.
The maximum peak values of bridge deck acceleration calculated along each track
shall not exceed the following design values:
(a) 𝛾 bt for ballasted track;
(b) 𝛾 df for direct fastened tracks with track and structural elements designed for
high-speed traffic;

SPAN: L = 30 M
Ref. Freq.: f = 3 Hz
Vertical displacements Ref Mass.: 25.000 kg/m
22
20 Thalys
18
Displacements (mm)

16 Ice-2
14
12
Eurostar
10
8
ETR-Y
6
4
2 TALGO_AV
0
16 18 20 22 24 26 28 30 32 34 36 38 40
Speed/f (m)

Vertical accelerations
6

5 Thalys
Acceleration (m/s2)

4 Ice-2

3 Eurostar

2 ETR-Y
1
TALGO_AV
0
16 18 20 22 24 26 28 30 32 34 36 38 40
Speed/f (m)

Figure 2.21 Dynamic amplification of deck deformation and acceleration.


50 2 Track for High-Speed Bridges

for all members supporting the track considering frequencies (including considera-
tion of associated mode shapes) up to the greater of:

(a) 30 Hz;
(b) 1.5 times the frequency of the fundamental mode of vibration of the member
being considered;
(c) the frequency of the third mode of vibration of the member.

Recommended values, considering a safety coefficient of 2, are the following:

𝛾bt = 3.5 m∕s2


𝛾df = 5 m∕s2 .

The dynamic analysis shall be undertaken using characteristic values of the load-
ing from the Real Trains specified. The selection of Real Trains shall consider each
permitted or envisaged train formation for every type of high-speed train permitted
or envisaged to use the structure at speeds over 200 km/h. In any case, the individual
project may specify the characteristic axle loads, spacings, and total number of axles
or train length, for each configuration of each required Real Train.
In some cases, a dynamic analysis is required for a Maximum Line Speed at the
Site less than 200 km/h (this and other aspects of dynamic analysis will be presented
in detail in Chapter 5).
Additionally, in the case of lines integrated into de trans-European rail system
network, these checks must be made for the 10 trains constituting the universal
High Speed Load Model (HSLM-A) defined by the ERRI D214 committee, collected
by the Eurocodes and the European Technical specifications for interoperability [8].
The HSLM-A covers the effects of existing and future high-speed trains that fulfil
the requirements established in Annex E of EN 1991-2.
On the other hand, dynamic acceleration calculations generally are made with
the assumption of track and wheels without irregularities. To quantify the effects
of these irregularities the maximum calculated acceleration must be multiplied
by (1 + 𝜑′′ ) for track with standard maintenance, or (1 + 0,5𝜑′′ ) for carefully
maintained track. The following expression is recommended for the computation
of 𝜑′′ :
[ ( )2 ( ) ( L )2 ]
v L
− Φ n0 LΦ − Φ
𝜌′′ = 0.56e 10 + 0.5 − 1 e 10
22 80
where

v is the Maximum Permitted Vehicle Speed (m/s)


n0 is the first natural bending frequency of the bridge loaded by permanent actions
(Hz)
LΦ is the determinant length (m) in accordance with 6.4.5.3 of EN 1991-2.

Considering that if v > 22, v = 22 will be taken, and 𝜑′′ = 0 if the result of the
computation is negative.
As abovementioned, more details will be presented in Chapter 5.
2.4 SLS Related to the Track 51

Figure 2.22 Definition of deck twist (source [10]). s

3m

Deck Twist Deck twist limits ensure that the four-wheel contact points of a bogie
(or of a two-axle vehicle) are not too far from a plane, assuring that rail-wheel
contact forces are controlled. In other case, the risk of derailment could be
increased.
The twist of the bridge deck shall be calculated considering the characteristic val-
ues of Load Model 71 as well as SW/0 or SW/2 as appropriate multiplied by the
dynamic coefficient Φ and the classification coefficient 𝛼 and Load Model HSLM
including centrifugal effects, all in accordance with EN1991-2, 6.
Twist shall be checked on the approach to the bridge, across the bridge, and for
the departure from the bridge.
The maximum twist t (mm/3m) of a track gauge s (m) of 1.435 m measured over
a length of 3 m (Figure 2.22) should not exceed the values given in Table 2.3.
The recommended values for the set of t are:
t1 = 4.5
t2 = 3.0
t3 = 1.5
Values for a track with a different gauge may be defined, but the following criterion
could be adopted:
t1 = 4.5 𝛽
t2 = 3.0 𝛽
t3 = 1.5 𝛽
where 𝛽 = 1.78r 2 /(r + c)2 and c = 0.5 m. Justification of this value is included in [9].

Vertical Deformation of the Deck Vertical deflection of the deck must be limited to
ensure acceptable vertical track radii and generally robust structures.
EN 1990 Annex A2 [10] establishes as general limitation, for all structure con-
figurations loaded with the classified characteristic vertical loading in accordance

Table 2.3 Limiting values of deck twist.

Speed range Maximum twist


V (km/h) t (mm/3m)

V ≤ 120 t ≤ t1
120 < V ≤ 200 t ≤ t2
V > 200 t ≤ t3
52 2 Track for High-Speed Bridges

with EN 1991-2, 6.3.2 (and where required classified SW/0 and SW/2 in accordance
with EN 1991-2, 6.3.3) the maximum total vertical deflection measured along any
track due to rail traffic actions should not exceed L/600.
Additional requirements for limiting vertical deformation (that can be different
for ballasted and ballastless bridges) may be specified. In the case of ballastless
bridges, it must be considered that, in the absence of ballast, the possibilities of
long-term correction of track geometry are limited to the adjustment capacity of the
fastening system. Thus, long-term deformation of the deck can affect track geometry
quality and should be added to deformations due to traffic load in the verification
abovementioned. The final goal is to assure that track geometry quality can be
maintained in any circumstances including long-term deformations of the deck.
On the other hand, the vertical deflection of the end of the deck beyond bearings
must be limited:

(a) to avoid destabilising the track in the case of ballasted track;


(b) to limit uplift forces on rail fastening systems and limit additional rail stresses
(to avoid premature failures, on both ballasted and ballastless bridges).

In this sense, EN 1990 Annex A2 indicates that limitations on the rotations of


ballasted bridge deck ends are implicit in Section 6.5.4. of EN 1991-2 , dedicated to
combined response of structure and track to variable actions, but that the require-
ments for ballastless structures may be specified in the National Annex.
For its part, EN 1991-2 indicates that for directly fastened rails the uplift forces
(under vertical traffic loads) on rail supports and fastening systems shall be checked
against the relevant limit state (including fatigue). A specific analysis, considering
performance characteristics of the rail supports and fastening systems, will be
necessary.
EN 1991-2 also indicates that the vertical displacement of the upper surface of a
deck relative to the adjacent construction (abutment or another deck) due to variable
actions shall not exceed the following values:

– 3 mm for a Maximum Line Speed at the Site of up to 160 km/h;


– 2 mm for a Maximum Line Speed at the Site over 160 km/h.

The criteria for the limit values will be determined by the tensile forces in the
fasteners. In this sense, the German standard for slab tracks indicates that the max-
imum traction in the fasteners on both sides of the joint must not exceed the value
that produces the decompression of the rail pad, a value that depends on the charac-
teristics of the fastening system and can be determined by standardised tightening
force test [11–13].
UIC leaflet 776-3R [14] indicates a design criterion regarding the limitation of
stresses in the rail, indicating that when there are direct fasteners on both sides of
the joint, the variation of angle and lifting should be limited to not exceed a stress
of the order of 80 N/mm2 in the rail. This stress should be calculated considering the
real position and stiffness of the fasteners. For rail type UIC 60 E1, this maximum
stress is equivalent to a maximum bending moment of 26.8 kNm.
2.4 SLS Related to the Track 53

From the above, to determine the maximum rotations and vertical relative dis-
placements the following criteria should be considered:
– comply with the explicit requirements indicated by standards;
– limit the bending moment in rails to assure that a stress of 80 N/mm2 is not
exceeded, as indicated by UIC leaflet 776-3R [14];
– limit the tensile force on the fasteners so that the clamping force is not exceeded,
determined in a standardised test, as indicated by the German standard.
Finally, rotation of the ends of each deck should be verified in the presence of rail
expansion devices, switches, crossings, etc. to limit angular discontinuity.

Transverse Deformation and Vibration of the Deck Transverse deformation and vibra-
tion of the deck must be limited to ensure acceptable horizontal track radii and to
avoid dynamic interaction with the vehicle.
According to EN 1990 Annex A2, the transverse deflection at the top of the deck
should be limited to ensure:
– a horizontal angle of rotation of the end of a deck about a vertical axis not greater
than the values given in Table 2.4, or;
– the change of radius of the track across a deck is not greater than the values in
Table 2.4, or;
– at the end of a deck the differential transverse deflection between the deck and
adjacent track formation or between adjacent decks does not exceed the specified
value.
Again, as in the case of vertical deformation, in the case of ballastless track the
effect of long-term deformation should be considered, if the lateral adjustment
capacity of the fastening system is not enough to correct these lateral deformations.
On the other hand, lateral deformations can influence the dynamics of vehicles.
As in the case of vertical deformations, in some viaducts an analysis of the dynamic
effects of lateral deformations may be necessary.
Standards, as EN 1990 Annex A2, establish a minimum value for the first natural
frequency of lateral vibration of a span of 1.2 Hz. This limit was proposed in by the
ERRI D181 Committee [15], dedicated to the analysis of lateral forces laterals on

Table 2.4 Maximum horizontal rotation and maximum change of radius of


curvature (source [10]).

Maximum change of radius


of curvature (m)
Speed range Maximum horizontal Single Multi-deck
V (km/h) rotation (rad) deck bridge

V ≤ 120 𝛼1 r1 r4
20 < V ≤ 200 𝛼2 r2 r5
V > 200 𝛼3 r3 r6
54 2 Track for High-Speed Bridges

railway bridges, and was justified in order to avoid lateral resonance phenomena in
railway vehicles circulating on the bridge, considering that, in general, the lateral
vibration frequencies of railway vehicles are below 1.0 Hz.

Note 1: The change of the radius of curvature may be determined using:


L2
R=
8 • 𝛿h
Note 2: The transverse deformation includes the deformation of the bridge deck and
the substructure (including piers, piles, and foundations).
Note 3: The values for the set of 𝛼 and r i may be defined in the National Annex. The
recommended values are:

𝛼1 = 0.0035; 𝛼2 = 0.0020; 𝛼3 = 0.0015;

r1 = 1700; r2 = 6000; r3 = 14 000;

r4 = 3500; r5 = 9500; r6 = 17 500.

However, it should be emphasised that this limitation refers to ‘the first natural
frequency of lateral vibration of one span’, which can be interpreted as the lateral
deformation of the deck assuming that the piers represent fixed points in lateral
direction. However, this assumption cannot be realistic in some cases, as, for
instance, the case of long viaducts with high piers. In this case, the lateral deforma-
tions occurring during a train pass-by can be significant and the natural frequencies
of the first mode of vibration of the deck can be very low. It is not clear whether the
required verifications must be applied to spans considered independently, to several
successive spans, or to the whole viaduct.
Standards do not propose any analysis methodology to assess this situation and
validate the design of this kind of viaducts. In some special cases, the consideration
of lateral dynamics for the response of railway vehicles on bridges will require the use
of three-dimensional models, including degrees of freedom for lateral displacement,
rolling, and yawing in the vehicles [16, 17].

Longitudinal Displacement of the Deck Due to the longitudinal interaction between


track and structure, the presence of railway bridges may generate added stress on
continuous rails and relative displacements between track and deck or substructure.
Thus, an analysis of these phenomena is needed from the structure design phase for
ensuring proper system behaviour.
UIC Leaflet 774-3R [18] and Eurocode EN 1991-2 [1] state methodologies for the
analysis of track-deck interaction, as well as actions to be considered and limit values
both for stress in rail and displacements. The following limits should be fulfilled:
Due to traction and braking 𝛿 B (mm) shall not exceed the following values:

– 5 mm for continuous welded rails without rail expansion devices or with a rail
expansion device at one end of the deck;
– 30 mm for rail expansion devices at both ends of the deck where the ballast is
continuous at the ends of the deck;
2.4 SLS Related to the Track 55

– movements exceeding 30 mm shall only be permitted where the ballast is provided


with a movement gap and rail expansion devices provided.

where 𝛿 B (mm) is:

– the relative longitudinal displacement between the end of a deck and the adjacent
abutment or;
the relative longitudinal displacement between two consecutive decks.

For vertical traffic actions, up to two tracks loaded with load model LM 71 (and
where required SW/0) 𝛿 H (mm) shall not exceed the following values:

– 8 mm when the combined behaviour of structure and track is considered (valid


where there is only one or no expansion device per deck);
– 10 mm when the combined behaviour of the structure and track is neglected.

where 𝛿 H (mm) is:

– the longitudinal displacement of the upper surface of the deck at the end of a deck
due to deformation of the deck.

Although these phenomena may be analysed theoretically, these reference


documents approach interaction computations by means of numerical models that
idealise the behaviour of all elements and actions involved to obtain stresses and
displacements.
However, some special issues not addressed in current standard should be consid-
ered, as:

– combination of actions: comparison of simplified combination methods versus


step-by-step computation methods;
– considerations on deformations due to shrinkage and creep in concrete
pre-stressed decks;
– special features of interaction with ballastless track;
– track–structure interaction under seismic conditions;
– application of tailored methodologies to studies on special geometries.

All these aspects will be explained in detail in Chapter 5.

2.4.3.2 Comfort Criteria


Criteria for ride quality are related to the acceleration levels and the associated
frequency ranges experienced by travellers inside the coach during travel on the
approach to, passage over, and departure from the bridge.
The expert ERRI committees carried out many tests and simulations to determine
the values of acceleration associated with different levels of passenger comfort. As
a result of these analysis, established general limits for deck deformation to ensure
that acceleration limits are fulfilled.
The methods adopted by the ERRI committees took full account of complex mod-
elling vehicle/structure, simplified modelling of track behaviour and took account
of the characteristics of the primary and secondary suspensions of the vehicle.
56 2 Track for High-Speed Bridges

Table 2.5 Recommended levels of comfort [10].

Vertical acceleration
Level of comfort inside the coach (m/s2 )

Very good 1.0


Good 1.3
Acceptable 2.0

The levels of comfort and associated limiting values may be defined for any indi-
vidual project, but recommended levels of comfort are given by EN 1990 Annex A2
(see Table 2.5).
As abovementioned, to limit vertical vehicle acceleration to the values given in
Table 2.5, limit values for the maximum permissible vertical deflection along the
centre line of the track of railway bridges were established as a function of the span
length, the train speed, the number of spans, and the configuration of the bridge
(simply supported beam, continuous beam). These values can be determined
through Figure 2.23.
The values in Figure 2.23 are given for a succession of simply supported beams
with three or more spans. For a bridge comprising either a single span or a succes-
sion of two simply supported beams or two continuous spans the values should be
multiplied by 0.7. For continuous beams with three or more spans the values should
be multiplied by 0.9.
It must be considered that the limiting values are given for an acceleration
inside the coach of 1.0 m/s2 which may be taken as providing a ‘very good’ level of

3.000
V=
350
2.500
V=
V= 300
280
2.000 V=
250
V=
V= 220
L/δ

1.500 200
V=
160
1.000
V=
120

500

0
0 10 20 30 40 50 60 70 80 90 100 110 120
L (m)

Figure 2.23 Maximum permissible vertical deflection for railway bridges with 3 or more
successive simply supported spans corresponding to a permissible vertical acceleration of
1 m/s2 in a coach for speed V (km/h) (source [10]).
References 57

comfort. For other levels of comfort and associated maximum permissible vertical
accelerations the values may be divided by the maximum permissible vertical
acceleration in m/s2 .
To be coherent with the computation done by ERRI D190 committee, the vertical
deflections should be determined with Load Model 71 multiplied by the factor Φ and
with the value of 𝛼 = 1, in accordance with EN1991-2, and, for bridges with two or
more tracks, only one track should be loaded.
It must be pointed out that the values of L/𝛿 given in Figure 2.23 are valid for span
lengths up to 120 m and for a speed up to 350 km/h. For longer spans or higher speeds
a special analysis should be necessary. Similarly, for exceptional structures, e.g. con-
tinuous beams with widely varying span lengths or spans with wide variations in
stiffness, a specific dynamic calculation should be also necessary.
The limits established for vertical deflection can be considered as too conservative
in some structures. In this case, as for the structures not covered by these require-
ments, complex analyses following the principles adopted by the ERRI D190
committee can be carried out subject to the agreement of the corresponding
infrastructure manager. To be able to justify a given design, these analyses must
consider all the types of actual trains proposed for operating the lines, as well as
others envisaged to be operated in the future.
It must be pointed out that the limits proposed by the ERRI D190 committee are
related only to comfort. The compliance with these limits favours compliance with
the limits of acceleration of the bridge deck and other dynamic aspects related to
security. However, compliance with these limits should not be considered as a guar-
antee that the safety criteria linked to resonance are met. The review of work of
ERRI D190 committee indicates that, for certain bridges (for example continuous
beams of length less than 40 m with a single loaded track) the deck acceleration
limit of 3.5 m/s2 could be exceeded if the travellers’ comfort criteria were consid-
ered to govern the design. Likewise, it is likely that the bridges meeting the safety
criteria defined in this report have acceptable characteristics for traveller’s comfort.
However, compliance with the criteria related to vertical acceleration of the deck
cannot be considered a guarantee that the travellers’ comfort criteria will be satisfied.

References

1 European Committee for Standardization (CEN). Eurocode 1 EN 1991-2. Actions


on structures – Part 2: Traffic loads on Bridges; 2003.
2 Romo, E., Nasarre, J., Lozano, A. et al. (2014). High Speed Railway Track Techni-
cal Options. Suitability Assessment Guide. UIC – High Speed Department.
3 Union Internationale des Chemins de Fer (UIC) “Vertical Elasticity of Ballastless
Tracks”. Project I/03/U/283. Paris. 2008.
4 European Committee for Standardization (CEN). EN 13848-1. Railway appli-
cations – Track – Track geometry quality. Part 1: Characterization of track
geometry. 2019.
58 2 Track for High-Speed Bridges

5 European Committee for Standardization (CEN). EN 13848-5. Railway applica-


tions – Track – Track geometry quality. Part 5: Geometric quality levels. 2017.
6 Union Internationale des Chemins de Fer (1994). UIC Leaflet 776-1R. Loads to
be considered in railway bridge design.
7 European Rail Research Institute (ERRI). ERRI D214 committee. Bridges for
speeds over 200 km/h. Final report. European Rail Research Institute. 1998.
8 European Commission. Technical specifications for interoperability relating
to the ‘infrastructure’ subsystem of the rail system in the European Union.
Commission Regulation (EU) No 1299/2014 of 18 November 2014.
9 Nasarre, J. “Serviceability limit states in track on railway bridges”, Railway
Bridges, Revista de obras públicas/June 2004/n∘ 3.445, pp 65-74. 2004.
10 European Committee for Standardization (CEN). Eurocode 0 EN 1990:2002/A1:
Basis of structural design. Annex A2: Application for bridges; 2005.
11 European Committee for Standardization (CEN). EN 13146-7. Railway appli-
cations – Track – Test methods for fastening systems. Part 7: Determination of
clamping force. 2012.
12 Centro de Publicaciones, Ministerio de Fomento. “Documentos complementar-
ios no contradictorios para la aplicación de los Eurocódigos para el cálculo de
puentes de Ferrocarril”. Dirección General de Ferrocarriles. 2014.
13 Goicolea, J.M., Cuadrado, M., Viñolas, J. et al. (2014), Estudio del compor-
tamiento a medio y largo plazo de las estructuras ferroviarias de balasto y placa.
Centro de Publicaciones, Ministerio de Fomento, CEDEX. 2013.
14 Union Internationale des Chemins de Fer (1989). UIC Leaflet 776-3R. Deforma-
tion of bridges.
15 European Rail Research Institute (ERRI). ERRI D181 committee. Lateral forces
in railway bridges. Rapport 6. 1996.
16 Cuadrado, M., González, P., Goicolea, J.M. et al., Analysis of lateral displace-
ments in large railway viaducts under traffic loads. Impact on ride safety and
passenger comfort, (2008), 8th World Congress on Railway Research. Seoul 2008.
17 Goicolea, J. and Antolín, P. (2012, 2012). The dynamics of high-speed rail-
way bridges: a review of design issues and new research for lateral dynamics.
International Journal of Railway Technology 1: 27–55.
18 Union Internationale des Chemins de Fer. UIC Leaflet 774-3R. Track-bridge
interaction. Recommendations for calculations. 2001.
61

Conceptual Design of High-Speed Railway Bridges


José Romo

3.1 Introduction
The first question that can be asked when designing a bridge for a high-speed railway
line is: What’s so special about a high-speed rail bridge?
Railways bridges have to meet strong structural requirements. The live loads in
those bridges are substantially higher than those of road bridges.
In the case of bridges for high-speed lines, the mass and speed of the trains interact
with the structure resulting in significant dynamic excitation of the structure, with
the consequent danger of resonance. Therefore, high stiffness is required to control
vertical acceleration to ensure the comfort of the passengers as well as the safety of
the trains. For this reason, the vertical and transverse deformations of the deck must
be controlled.
Other special conditions such as high braking forces or the interaction between
track and bridge deck are specific in HSR bridges as the use of continuous welded
rail CWR is indispensable to maintain the regularity of rolling [1].
Good design should meet structural and functional requirements with a high
visual quality. To achieve that goal, the design of HSR bridges has to assure a sub-
stantial vertical structural stiffness without using massive elements. Furthermore,
high-speed railways lines are the best example of the high technology of the 21st
century; HSR bridge’s design should meet public expectation of those especial
elements.
Throughout this chapter, first the specific conditions of this type of bridges are
presented as well as the implications they have on the design of the structure. After
that, the standard design situations are then presented and the different structural
types associated with these situations are explained. Finally, a description is made
and a design methodology of the specific structural elements of this type of bridges
is presented.

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
62 3 Conceptual Design of High-Speed Railway Bridges

3.2 Structural and Functional Specific Requirements for


High-Speed Railway Bridges

3.2.1 Introduction
– The specific aspects that influence the design of bridges for HSR are as follows.
Vertical stiffness to control the vertical acceleration produced by the coupling of
the vibration of the mass of the structure and the trains.
– Vertical stiffness to control the relative rotation at expansion joints.
– Longitudinal fixation of the deck to transmit the huge braking forces to the
foundation.
Special provisions to limit the stresses in the rails due to the track to bridge deck
interaction.
Although all of these aspects are dealt with extensively in other chapters, they are
briefly listed here for a better understanding of the conceptual design of this type of
structures.
In Sections 3.2.2–3.2.6 all those aspects are analysed as well as its implications in
the conceptual design of this type of bridges.

3.2.2 Control of Vertical Acceleration


The high speed of the passing trains entails the repetition of dynamic loads on the
bridge. The coupling of the frequency of the vertical dynamic load with the natural
frequency of the deck may lead to high vertical acceleration due to resonance. Those
vertical accelerations could cause a lack of comfort for the passengers, as well as loss
of the contact of the track with the bridge due to a negative vertical acceleration.
For that, the vertical stiffness has to be substantially higher than the typical one in
highway bridges. This implies that the decks are in general heavier and have a bigger
depth and greater self-weight, with the direct consequences in terms of loads on the
substructure (see Section 2.4).

3.2.3 Rotation at Expansion Joints


To assure the security of the passing trains, it is necessary to limit the relative rotation
at the expansion joints. For that, the deck has to have enough vertical stiffness to
meet those limits. As in the previous case, this condition means that the deck must
have a higher bending rigidity than other types of bridges and therefore the deck
must have more depth and more stiffness (see Section 2.4).

3.2.4 Horizontal Braking and Traction Forces and Relative Movements


Between Deck and Infrastructure
The braking and traction forces in HSL bridges are as high as 12 MN in long viaducts
with continuous deck (see Figure 3.1).
3.2 Structural and Functional Specific Requirements for High-Speed Railway Bridges 63

14.00

12.00
Force in fixed point (mn)

10.00

8.00

6.00

4.00

2.00

0.00
0 100 200 300 400 500
Viaduct length (m)

Figure 3.1 Braking and traction force vs deck length.

Another condition to assure the safety of the trains is the maximum relative defor-
mation between deck and infrastructure. The value of this relative movement is as
low as 5 mm when braking forces arc, which in practice entails that the deck has to
be longitudinally fixed to one rigid point of the infrastructure, normally one of the
abutments. That fixed point has to pass the large braking and traction forces from
the deck to the infrastructure. It is therefore necessary to design at least one element
of the substructure, either an abutment or an intermediate support, capable of trans-
mitting these large horizontal loads in the longitudinal direction (see Section 2.4).

3.2.5 Track-Bridge Deck Interaction


Due to continuity of the rail and the high braking forces and the differential temper-
ature between the track and the structure, it is necessary to evaluate the additional
stresses in rails and in case to install an expansion joint in the rails to limit those
stresses. Depending on the length of the bridge, this aspect is critical, as it influ-
ences the position of the fixed point, discussed above, and therefore the strategy for
resisting the longitudinal loads of the deck.

3.2.6 Expansion Joints


Track expansion joints are always a maintenance problem and should therefore be
avoided as much as possible. If expansion joints are used, the value of their stroke
significantly influences the positions of the fixed points and thus the longitudinal
configuration of the bridge.
At the beginning of high speed, the maximum track joint stroke was only 600 mm,
which considerably limited the length of continuous sections. Today, this value is
up to 1200 mm, which has made it possible to achieve much longer continuous deck
lengths.
The expansion joints of the structure must be perpendicular to the track axis to
ensure the safety of rail traffic. When the crossing is skewed, either the deck must
be made longer so that a joint can be made perpendicular to the tracks, or the deck
and the abutment must be divided to offset the abutment so that each section of the
structure joint is perpendicular to the track above it [2].
64 3 Conceptual Design of High-Speed Railway Bridges

3.3 Longitudinal Design Strategies


3.3.1 General Concepts
Of the above-mentioned factors, the track–structure interaction and the limitations
on differential length movements between the deck and the infrastructure are
the factors that most influence and differentiate this type of bridges from railroad
bridges. In HSR bridges the deck and the infrastructure must be rigidly connected
against longitudinal, as well as against transversal movements. From a longitudinal
point of view, there are several ways to resolve this issue.
The first consists of fixing the deck to an abutment, leaving the rest of the deck’s
supports in piers and on the opposite abutment longitudinally free (see Figure 3.2).
This solution is feasible for continuous decks of short lengths (less than 100 m).
A second option, also feasible for continuous decks, consists of materialising the
longitudinal fixed point in an intermediate element with large longitudinal rigidity
(Figure 3.3). This is the typical case of continuous viaducts with a maximum length
of 2000 m (1000 m from the central fixed point to each expansion joint).
The third option is only valid for simply supported decks. In this case each deck
section is fixed longitudinally to a pier, while it is free in that direction in the other
pier it rests on (Figure 3.4).
In the case of ultra-long and continuous viaducts, with lengths of more than 2 km,
it will be necessary to have more than one longitudinal fixed point and an interme-
diate rail expansion joint (Figure 3.5).
An alternative to this solution is the use of a simply supported span, which is
located between two continuous lengthy sections in which two track expansion

Expansion joint Fixed point


92.00
16.00 20.00 20.00 20.00 16.00

Figure 3.2 Short viaduct with a longitudinal fixed point in an abutment.

Expansion joint Expansion joint


L < 1000 m L < 1000 m
Fixed point

Figure 3.3 Long viaduct with a longitudinal fixed point in a central pier.

Structural Structural Structural


expansion joint expansion joint expansion joint

Fixed Longitudinally Fixed Longitudinally Fixed Longitudinally


guided guided guided

Tranversally Tranversally Tranversally


guided Free Free Free
guided guided

Figure 3.4 Bearings layout for simply supported spans.


3.3 Longitudinal Design Strategies 65

Expansion joint Expansion joint Expansion joint Expansion joint


L < 1000 m L < 1000 m L < 1000 m L < 1000 m

Figure 3.5 Ultra-long and continuous viaducts with two intermediate fixed points.

joints are placed consecutively, in order to reduce their movement and thus make it
feasible to have two consecutive structures of more than 1 km each. This solution,
which is perfectly valid from a structural point of view, has the disadvantage for
rail traffic of having two consecutive rail joints, which are always delicate elements
from perspective of maintenance and train traffic safety.
The practical consequence of those main possibilities described depends on
whether the track is ballasted or not. The following sections show the possible
scenarios for each of these track types.
For bridges in seismic zones, the design for braking loads, and therefore the need
for a fixed point, has to be reconciled with the need to dissipate energy. This means
that longitudinal dampers have to be provided on these bridges to reduce the forces
caused by an earthquake [1, 3]. Those aspects are described in Section 3.7

3.3.2 Ballasted Track


In the case of continuous decks if the distance between the fixed point and the far-
thest point of the deck is less than 90–100 m expansion joints in the track are not
required. In the case of using one expansion joint in the track with 1200 m of stroke,
the maximum distance between the fix point of the structure and the expansion joint
of the track has to be around 1000 m, for concrete bridges. That means that with
the fixed point in one abutment, the maximum length of the bridge with one single
expansion joint in the other abutment is 1000 m; meanwhile if the fixed point is in
the centre, the maximum length of the bridge could be 2000 m having two expansion
joints at the abutments for concrete bridges (see Figure 3.6) and around 3000 m for
composite (steel and concrete) decks.

100 m < L

Fixed point Structural


at abutment expansion joint

Track
100 m < L < 1000 m expansion joint

Fixed point Structural


at abutment expansion joint

Track 1000 m < L < 2000 m Track


expansion joint Fixed point at centre expansion joint

Structural Structural
expansion joint expansion joint

Figure 3.6 Schemes configuration of continuous concrete decks with different length.
66 3 Conceptual Design of High-Speed Railway Bridges

All those requirements have to be considered during the conceptual phase. They
are very strong from a structural point of view and have a significant impact in the
visual aspect of the bridge. It is extremely easy to end up with a heavy structure if the
design does not take care of the visual aspect of the bridge. The balanced equilibrium
of all those aspects requires a careful design. In the following sections a discussion
of the different design situation is presented.

3.3.3 Ballastless Track


In the case of ballastless track, the longitudinal connection between rail and
structure depends to a large extent on the stiffness of the elastomer interposed
between the deck and the slab, the rail to sleeper, sleeper to slab connection, etc.
(see [4]). Therefore, the criteria indicated for ballasted track are not completely
applicable to ballastless track.

3.3.4 Actions to be Considered at the Fixed Point


The horizontal forces to be taken into account for the design of the fixed points are
the braking and acceleration forces together with friction of bearings, longitudinal
wind and temperature, with their corresponding combination coefficients.
Friction force at bearings (see EN-1337-1). In addition to the friction values to be
adopted, this standard indicates the possible reduction of forces as the number of
bearings increases.

3.4 Design Situation of High-Speed Railway Bridges


Definitely the strong structural requirements, in terms of loads and stiffness, do not
allow the designer to speculate too much with the shape of the structure. In this type
of bridges the geometrical form has to follow the structural logic. Here there is no
room for formalism without structural sense. It is clear that this type of bridges has to
be robust, but does not mean that the structure has to appear extremely heavy. Many
of the bridges shown in the Appendix: Built Bridges are examples of well-designed
HSR bridges. They combine structural and functional requirements with high visual
quality.
The purity of the structural shapes has to accommodate to the specific site con-
straints, and therefore the strategy has to be different depending on the nature of
the obstacle to be spanned.
In Section 3.4, different approaches to the diverse design situations are presented.
Among the high variety of possible scenarios, some of the most common situations
are here presented:
– short crossings at low level,
– long and low viaducts,
– long and high viaducts,
– long span HSB.
3.4 Design Situation of High-Speed Railway Bridges 67

In all the cases, a coherent approach with the structural performance as a clear
driver of the solutions has to be used.

3.4.1 Short Crossing at Low Level


That is for instance the crossing of a HSR over a highway or a short obstacle. In this
case, the length to be spanned plays the most important role to define the structural
solution.
In general it will be necessary to decrease to a minimum the deck depth to have
enough vertical clearance under the bridge. U cross sections could be adequate
when the span has a maximum of 30–35 m. For longer spans, other solutions such
as arches, ‘inverted frames’ or trusses could be used. Here, as the railway is normally
closed to the ground, it is important to avoid the massiveness and the designer has
to shape properly the structural members to improve the visual quality of the bridge.
In Figure 3.7 an example of a very skew crossing of a HSRL over a busy highway is
shown [5].There the use of a truss structure in which the design tried to reduce the
visual mass of the elements was one of the goals of this structure.
The proper shaping of the structural elements as well as the elimination of redun-
dant members contributes to the visual appearance of the bridge.

3.4.2 Long Structures


This is one on the most frequent situation in HSR Lines, since the radii of curvature
in plan are very large and consequently it is complex to adapt to the terrain when
the orography is varied.

14.40
7.60

Figure 3.7 Riudellots Viaduct, Spain (Source: FHECOR).


68 3 Conceptual Design of High-Speed Railway Bridges

Long viaducts are required for crossing of long valleys or just the design of the
railways just a few meters over the ground surface to avoid placing the line over a
filling.

3.4.2.1 Low Profile


In some cases, especially in horizontal landscapes, the railway is designed just a few
meters above the ground. From the point of view of the line itself, the low profile
entails low filling which has the benefit of having small settlements. But at the same
time, it is normally necessary to have long viaducts to have the required transversal
transparency under the line.
It should be known, that the normal depth to span ratio for a continuous concrete
prestressed viaduct for a HSL is around L/16 for cast in situ solutions (Figures 3.8
and 3.9) and of L/12 for precast elements (Figure 3.10).
If the vertical clearance under the bridge usually required is in the range of
5.50 m, the alignment should be around 8.00 m over the ground level. In the case
of extremely low viaducts, U cross sections it is the best solution as it allows the
maximum free vertical clearance under the deck. In all the cases the lateral faces
required a special treatment as they are the most visual element of the bridge.
In the case of steels structure a light folding of the webs, increased the buckling
capacity of the steel plates as well as the appearance of the structure. In the case of
concrete structure is also possible to improve the lateral appearance by creating two
planes that could break the visual ‘wall effect’ (Figure 3.11).

~L/17 to ~L/14

Figure 3.8 Example of in situ deck for 20 m of span.

~L/17 to ~L14

Figure 3.9 Example of in situ deck for 30 m of span.

~L/12

Figure 3.10 Example of precast deck for 30 m of span.


3.4 Design Situation of High-Speed Railway Bridges 69

~L/10

Figure 3.11 Example of U section in concrete.

Figure 3.12 Monotonous bridge.

Figure 3.13 Continuous frames.

In all those cases, the repetition of the spans could result in a viaduct with a kind
of millipede’s appearance (Figure 3.12). Furthermore, it is very usual to increase
the vertical clearance under the deck by using short spans which increase the
monotonous view of the structure, which also creates a visual barrier when the
bridge is observed in a skew angle.
A possible way of improving the aesthetic of those ‘millipede’ bridges is increasing
the span by using a continuous frame structure that also decreases the number of
foundations (see Figure 3.13).
Other alternative when the designer wants a more openness under the bridge
could be the use of an ‘inverted frame’ system. In that case, the deck should be in
steel and the result is a more open solution that could be used for instance when the
line has to cross over a highway (see Figure 3.14).
From the structural point of view both solutions can fulfil the structural conditions
related to HSL bridges without problems.
In the first case, the frames have to be supported by pot bearings (Figure 3.15).
Those elements transmit the vertical forces to the foundation and allow the hori-
zontal movement to the fix point usually placed in one abutment.
In the case of the inverted frames, the deck and the upper part of the columns
transmit the vertical forces to the part of the supports under the deck by pot

Figure 3.14 Continuous inverted frames.

Bearing

Figure 3.15 Bearing and pier layout for continuous frame.


70 3 Conceptual Design of High-Speed Railway Bridges

Bearing

Figure 3.16 Bearing and pier layout for continuous inverted frames.

bearings. In the same way, the deck could be fixed monolithically to one abutment.
In Figure 3.16 a possible bearing layout in a standard pier is shown.
If those structures when the length of the deck is longer than 1000 m, a central
fixation is necessary as it was discussed previously. That central point attracts the
braking and traction forces, and therefore it is very difficult to resist that force by
the bending of the columns. There are several ways of designing those elements as
it can be seen in Figure 3.17. In all those cases special attention to the layout and the
connection of the different elements should be carried out to maintain the visual
quality of the structure but endowed with an appropriate control of the forces.
In some cases these long, low viaducts need a special, wider span to overcome a
special obstacle such as a motorway or river.
In such cases, it is both structurally and visually desirable for the deck to be con-
tinuous, although it will logically require an additional superstructure to bridge the
special span(s). The figure shows the studies carried out for the resolution of such a

2 pots
Pl – 14000

2 pots
Pl – 14000
2 neoprenes 2 neoprenes
400 × 400 400 × 400
2 neoprenes
400 × 400

30° 30°

3.0 0
0 3.0

4 cables 4 cables
2 × 4 cables
19T15 19T15
19T15
R1.
2.20
1.95

95

1.8 0
0 1.8

Figure 3.17 Solution types for intermediate fixed point.


3.4 Design Situation of High-Speed Railway Bridges 71

Figure 3.18 Study of a long crossing with a single main span.

type of crossing. In that case, a solution with a variable upward depth, a truss, and
an inverted frame was analysed (Figure 3.18).

3.4.3 High-Level Viaducts


The route of the high-speed railway has very wide radii of curvature in plan and
very low longitudinal slopes. This means that, in areas with a rough topography,
long crossings over relatively high valleys take place.
When it is possible a clean crossing with high piers could work also visually,
because the height of the structure over the valley reduced the visual mass of the
elements.
The high viaducts have a very noticeable influence on the environment. F. Leon-
hardt [6] indicates that from a landscape point of view the span length, in high and
long viaducts in U-shaped valleys, should be one and a half to two times the pier
height. In the case of railway bridges, it is rare to reach a maximum span of 60 m, so
the rule suggested by Leonhardt could only be applied to bridges with a maximum
pier height of approximately 40 m.
Figure 3.19 shows a typical solution where the length of the spans is very
similar to the height of the piles. It could be said that this solution is still visually
pleasing and represents an acceptable compromise between economy, construction,
structural behaviour, and aesthetics.
In the case of long structures when the length is over 1000 m the central anchor-
ing it is also necessary. Here due to height of the bridge, the design of the central

Figure 3.19 Continuous viaduct.


72 3 Conceptual Design of High-Speed Railway Bridges

Figure 3.20 Examples of central fixation.

Expansion joint Expansion joint Expansion joint

Figure 3.21 Gänsebachtal viaduct (sbp).

anchoring becomes more critical because of its visibility (Figure 3.20). Again, there
are some possibilities and when it is possible the use of one singular point of the
topography as the cross of a river or a higher intermediate point in the landscape
could be used to locate the central point [7].
The German railways have developed a solution for bridges with moderate pier
heights: 15–20 m, with a series of fixed points (approximately every 110 m) with
intermediate structural expansion joints between the sections, but without rail
expansion joints in the track [8]. An example of this solution is the Gänsebachtal
viaduct [9], shown in Figure 3.21. The bridge has a succession of flexible piers and
more rigid diapason piers that make a local fixed point.

3.4.4 Long Span Structures


In some cases, the structure has to jump over an important obstacle. In the case
of high-speed lines, structures with spans over 100 m could be considered as spe-
cial structures. In those cases, arches, truss, rigid cable-supported bridges or hybrid
structures are the structural types more adequate to combine stiffness and lightness.
In the following sub-chapters a discussion of the possible solutions is presented.

3.5 Structural Types


3.5.1 Straight Deck Solutions
This section includes the most commonly used structural solutions for HSR bridges.
For the most standard solutions, the basic dimensions as well as the weights and
their normal quantities are included.

3.5.1.1 Simply Supported Deck


The simply supported solutions logically have a structural expansion joint between
spans in the pier section. The great advantage of this solution is that it does not
require a rail expansion joint. Therefore, it is a solution with great advantages from
3.5 Structural Types 73

Figure 3.22 Simply supported launching beam system for 32 m prefabricated concrete
decks, China (Courtesy of China Railways).

the point of view of railway operation. The biggest disadvantage is its low structural
efficiency and the higher maintenance requirements for the bridge as there is a struc-
tural joint in every span.
It should also be noted that in a seismic zone this solution has the risk of loss
of support for the deck. Likewise, the lack of transverse continuity between spans
considerably increases the risk of rail traffic during the action of the earthquake.
This solution is the most widely used in some of the Southeast Asian countries.
Specifically, railway lines in China are being built with this type of solutions. In this
country, a complete prefabrication solution has been developed for a section that
also allows assembly from the previously built deck (Figure 3.22).
Also in Spain, isostatic solutions consisting of two prefabricated U girders for
spans of 30–35 installed with a launching beam are common.

3.5.1.2 Continuous Slab Concrete Decks


In the case of small spans, i.e. below 30–35 m, it is common to use continuous deck
solutions in prestressed concrete slabs. The prestressed concrete section consists of
a central core and lateral overhangs to complete the total width of the deck. The
central core will have longitudinal voids with a circular cross section or any other
suitable for concreting. Normally, these are solutions to be used when the level of
the railway in relation to the natural terrain is relatively low.
Figures 3.23 and 3.24 show a typical section of this type of viaducts (span of 32 m)
built on self-supporting formwork, span by span.
The classic parameters and average quantities of this type of bridges are
summarised in Table 3.1.
74 3 Conceptual Design of High-Speed Railway Bridges

14.00

2.80 4.20 4.20 2.80


2.35 2.35
1.76

0.20
Figure 3.23 Cross section slab solution.

A-1 P-1 P-2 P-3 P-4 P-5 P-6 P-7 P-8 P-9 P-10 P-11 P-12 A-2
402.00

25.00 32.00 32.00 32.00 32.00 32.00 32.00 32.00 32.00 32.00 32.00 32.00 25.00

Fixed point

Expansion
rail joint

Figure 3.24 Slab solution elevation.

Table 3.1 Summary of dimensions, quantities, and loads of slab concrete decks 14 m wide.

Span range 25 to 35 m
Total depth related to span length L L/14 to L/17
3 2
Concrete deck average equivalent thickness m /m 0.75 to 0.95
Average prestress weight in deck per surface deck unit kg/m2 18 to 30
2
Average steel rebars weight in deck per surface deck unit kg/m 80 to 105
Standard self-weight (kN/m) 260
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 430

If the height above the ground does not exceed 15 m, it is usual to build with
ground-supported formwork span by span (Figure 3.25).
For higher heights or if the terrain has a low bearing capacity, the construction
can be carried out by means of movable a scaffolding system supported by the bridge
piers.

3.5.1.3 Precast Beam Decks


Prefabricated beam decks are suitable in the low span range (15–40 m span) and
especially where lower traffic must be respected during construction or in the cases
where formwork bearing in the ground is not feasible.
The solution with isostatic beams can reach a span of up to 40 m, although it is
common for them to be as short as 35 m. Redundant solutions can reach up to 45 m,
although they are usually used up to 42 m span.
3.5 Structural Types 75

Next-construction Phase under


phase construction Built phase

L/5 Concrete L/5 Concrete


joint joint

Preparation of the
next falsework Falsework
phase

Figure 3.25 Example of slab-type deck with span-by-span construction with conventional
formwork.

In the case of redundant structures the common practice is to give continuity con-
necting adjacent spans by posttension bars. Normally, the cross section is formed by
two U-shaped beams located under the tracks and an in situ concrete slab built over
them (Figures 3.26 and 3.27).
Table 3.2 summarises the main dimensions, quantities, and weights of this type of
decks.

14.00
1.75 10.50 1.75

2.35 2.35
0.30
0.44

2.50

2.73 2.73

Figure 3.26 Standard solution with two U-shaped beams with structural continuity over
the piers.

A-1 P-1 P-2 P-3 A-2


126.00
27.00 36.00 36.00 27.00

Figure 3.27 Standard solution with two U-shaped beams (elevation).


76 3 Conceptual Design of High-Speed Railway Bridges

Table 3.2 Summary of dimensions, quantities, and loads of precast concrete decks 14 m
wide.

Span range (simple supported beams) 15 to 35 m


Span range (continuous deck) 15 to 40 m
Total beams depth related to span length L (simple supported beams) L/13.5
Total beams depth related to span length L (continuous deck) L/16
Concrete slab depth 0.30
Concrete deck average equivalent thickness m3 /m2 0.45 to 0.50
Average prestress weight in deck per surface deck unit kg/m2 11 to 18
3 3
Average steel rebars weight in deck per concrete m unit kg/m 130 to 210
Standard self-weight (kN/m) 180
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 350

14.00
1.75 10.50 1.75
2.35 2.35

0.30
2% 2%

3.00

2.50 4.50 4.50 2.50

Figure 3.28 Pre-fabricated solution.

By increasing the number of beams, it is possible to reduce the depth of the deck.
Figure 3.28 shows a section formed by three U-shaped beams.
The most common system for the assembly of this type of solutions is by means
of cranes (Figure 3.29). One of the advantages of this prefabricated solution, when
assembled with the help of cranes, is that it allows the independent construction of
each span in the order that is most convenient for the work.
An alternative to cranes is the use of a beam launcher. This solution is the most
reasonable possibility when the deck is at a very high level from the ground, or when
the prefabricated beams cannot be transported to the vertical of the area from which
they are to be lifted (Figure 3.30).
The continuity of the deck is carried out in the second phase, often by means of
post-tensioning, to make the connection section work as a pre-stressed joint with
limited crack opening. Figure 3.31 shows a typical pre-stressed connection in pre-
fabricated U-beams.
3.5 Structural Types 77

Figure 3.29 Beams’ delivery by means of cranes.

A-1 P-1 P-2 P-3 P-4 P-5 P-6 A-2

Figure 3.30 Beams’ launching beam system.

Precast 4Ø40 Precast


Beam depth 2.60 Slab
0.25

planks Prestressed planks


e = 0.06 rebars e = 0.06

4Ø40
0.06

Prestressed
Joint filled rebars
0.14 with grout
6Ø50
6Ø50 Prestressed
Prestressed rebars
rebars
6Ø50 6Ø50
Prestressed Prestressed
rebars rebars

Figure 3.31 Typical pre-stressed connection in prefabricated U-beams.


78 3 Conceptual Design of High-Speed Railway Bridges

14.00
7.00 7.00
2% 2%
0.20

0.40
4.40

4.26
0.55 0.55

0.30

77
°
3.35 0.90 5.50 0.90 3.35

Figure 3.32 Typical concrete box cross section.

3.5.1.4 Concrete Box Hollow Sections


Above 35–40 m, the most competitive solution is continuous concrete prestressed
decks with a box section. With this typology and constant depth deck, spans of up
to 80 m can be achieved, although it is usual that with spans longer than 70 m, the
most competitive solutions are those with variable deck depth, which can reach up
to 150 m (Figure 3.32).
In this type of sections, the width of the bottom of the section is usually 0.35 of the
section width. Web thicknesses vary from 0.45 to 0.60 m, and bottom slab thicknesses
from 0.30 to 0.50 m (see Figure 3.33).
The average parameters and quantities of the deck are summarised in Table 3.3.
These bridges are usually built by launching or by means of a movable scaffold-
ing system. The depth of railway bridges decks makes the launching technique
interesting in the case of decks with a constant hollowed box section, despite the
60.00 70.00
0.40

0.40
0.40

0.40

0.40

1.50
1.25
1.25

3.40
3.40
4.40

4.40
3.70

4.40

3.70
3.70

2.15
2.15

0.80
0.60
0.60

0.80
0.30

0.30

1.50 1.00
0.30

11.70 1.50 1.50 11.70

Figure 3.33 Typical concrete box longitudinal cross section.

Table 3.3 Summary of dimensions, quantities, and loads of concrete box decks 14 m wide.

Span range 40 to 70 m
Total depth related to span length L L/12.5 to L/16.5
Deck average equivalent thickness m3 /m2 0.75 to 0.90
Average prestress weight in deck per surface deck unit kg/m2 22 to 40
Average steel rebars weight in deck per surface deck unit kg/m2 140 to 150
Standard self-weight (kN/m) 280
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 450
3.5 Structural Types 79

magnitude of the loads to be pushed. Launching is usually competitive for bridges


longer than 400 m and with spans between 40 and 75 m (Figures 3.34 and 3.35).
The prestressing values given in Table 3.3 correspond to a conventional construc-
tion span by span. When the deck is pushed from one of the abutments, all the
sections of the deck will be subject to sagging and hanging bending moments as a
result of passing through different positions (span centre and pier supports). There-
fore, the amounts of pre-stressing required are higher (around 20% more) than those
indicated in Table 3.3.
Another very frequent alternative for the construction of deck is by means of a
movable scaffolding system. This solution is used when the height of the bridge
above the ground does not allow the use of conventional formwork and when the
length of the spans does not exceed approximately 60 m.
Two movable scaffolding systems (MSS) schemes are shown in Figure 3.36. The
upper part of the figure shows the case where the MSS is placed above the deck.
This is the usual situation when trying to solve a low viaduct with a short vertical
clearance under it. The lower part of the figure shows the most common case where
the MSS is located under the deck. Figure 3.37 shows the example of a MSS with the
structure over the deck.

Figure 3.34 Launching construction.

A-1 P-1 P-2 P-3 P-4 P-5 A-2

A-1 P-1 P-2 P-3 P-4 P-5 A-2


D-18
D-12 D-13 D-14 D-15 D-16 D-17

A-1 P-1 P-2 P-3 P-4 P-5 A-2


D-18
D-1 D-2 D-3 D-4 D-5 D-6 D-7 D-8 D-9 D-10 D-11 D-12 D-13 D-14 D-15 D-16 D-17

Figure 3.35 Launching construction.


80 3 Conceptual Design of High-Speed Railway Bridges

Figure 3.36 Construction by movable scaffolding system (MSS).

Figure 3.37 Construction of a concrete box section with an upper MSS.

3.5.1.5 Steel Beam Decks


Composite bridges with a double ‘I’ steel beam are used in the 40–70 m span range,
although they are more common in the 50–60 m range. Those decks have a lighter
self-weight than concrete structures, so they are usually an interesting solution in
those cases with high piers and/or bad foundation terrain [10] or when the bridge is
located in a high seismic area [11].
This solution is specially used in the range of spans over 65 m that are not reached
by concrete decks built with a movable scaffolding system (Figure 3.38).

14.00
1.75 10.50 1.75
2.35 2.35
0.22
0.41

3.00

6.00

Figure 3.38 Typical cross section of a double I steel beam.


3.5 Structural Types 81

Table 3.4 Summary of dimensions, quantities, and loads of steel double I girder 14 m wide
decks.

Span range 40 to 70 m
Total steel beam depth related to span length L L/14.5 to L/17
Total depth related to span length L L/13.5 to L/15
2
Deck average steel weight kg/m 270 to 300
Deck average equivalent thickness m3 /m2 0.40
3
Average steel rebars weight in deck per concrete volume kg/m 140 to 150
Standard self-weight (kN/m) 180
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 350

In the specific case of piles over 80 m high, the pushing of a deck with double ‘I’
steel beam is much safer than that of heavy concrete decks, so composite solutions
can be very competitive for those cases (Table 3.4).
Composite bridges with a double ‘I’ steel beams can be built as well span by span
using cranes if access is possible and the deck height is moderate (up to 35–40 m)
and by launching in other cases.

3.5.1.6 Steel Semi-through Decks


Another classic steel solution when there is a strict vertical clearance is the so-called
semi-through deck. In this case, the two steel beams are placed above the level
of the platform, freeing the space underneath from the main structural elements
(Figure 3.39).
Under the platform, a reinforced concrete slab is placed that distributes the loads
to steel crossbeams that in turn transmit the load to the longitudinal resistance

14.00

C
L

Figure 3.39 Example of semi-through deck with lateral hollow steel beams.
82 3 Conceptual Design of High-Speed Railway Bridges

Table 3.5 Summary of dimensions, quantities, and loads of steel semi-through 14 m wide
decks.

Span range 40 to 70 m
Total depth related to span length L L/10 to L/12
Deck average steel weight kg/m2 400 to 500
Concrete deck average equivalent thickness m3 /m2 0.40
Average steel rebars weight in deck per concrete volume kg/m3 250
Standard self-weight (kN/m) 215
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 385

system formed by the two main steel beams located at the edges of the platform.
The usual characteristics of this type of decks are summarised in Table 3.5.
One of the usual advantages of this solution is that it usually allows the push from
one of the abutments, avoiding the need for intermediate supports for the construc-
tion of the deck and minimising the effect on the obstacle to be spanned.

3.5.2 Truss Bridges


Steel truss bridges are a classic solution for railway bridges. Both in the case of strict
height bridges, as when it is a high viaduct, trusses are solutions that combine rigid-
ity with lightness.
Modern trusses try to follow structural logic to increase the structural effectivity
as well as the aesthetic that stems from the self-expression of the structure.
For long spans, the truss solutions tend to be continued and they must have a
variable depth, to better adapt to the bending moments law (Figure 3.40). Table 3.6
summarises the main dimensions, quantities, and weights of the truss solutions.
The trusses can be erected by launching or by crane erection if temporary supports
are possible and spans are limited. In special cases, it is also possible to assemble the
trusses by segments.

Figure 3.40 AP-7 Viaduct Almeria, Spain (Source: FHECOR).


3.5 Structural Types 83

Table 3.6 Summary of dimensions, quantities, and loads of steel truss decks.

Span range 50 to 240 m


Total depth related to span length L
• Constant depth
• Variable depth L/10 to L/13
– At mid span L/30 to L/25
– At supports L/12 to L/14
Deck average steel weight kg/m2
● Span length < 150 m 500 to 700
● Span length > 150 m 700 to 1000
Concrete deck average equivalent thickness m3 /m2 0.40
3
Average steel rebars weight in deck per concrete volume kg/m 250
Standard self-weight (kN/m) 220 to 280
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 390 to 450

3.5.3 Arch Bridges


3.5.3.1 Upper Deck Bridges
Arches are probably the main solution when the span length of the bridge is quite
long (see Figure 3.41). The concrete arches give the required stiffness with a limited
use of mass. Several examples around the globe show the proper use of this typol-
ogy for spanning lengths in the range of 200 to 350 m. Arches also have an intrinsic
beauty that makes them a safe choice from an aesthetic point of view.
In Table 3.7 a summary of the main dimensions of this type of bridges is presented.
As the arc span increases, the problems associated with transverse behaviour esca-
late becoming the most relevant aspects when designing the bridge.
In general, the deck for a two-way line is in the range of between 13 and 15 m
wide and, as there are, in general, approaching viaducts, it is not possible to use the
deck as a transversal stiffening element of the system. Therefore, it is important to
increase the transverse depth of the arch at its spring’s sections. For the longer spans,
one way to achieve this transverse stiffening is by widening and dividing the arch at
its springs into two parts (Figure 3.42).
The depth the deck is conditioned in this type of bridges by the span between
the supports on the columns resting on the arch. The distance between these

Figure 3.41 Example of upper deck concrete arch.


84 3 Conceptual Design of High-Speed Railway Bridges

Table 3.7 Summary of dimensions of upper deck arch bridges.

Span range 200 to 350 m


Span/raise 1/5.5 to 1/6.8
Depth of the arch related to span length L
At springs L/55 to L/70
At crown L/80 to L/90
Depth of the arch related to span length L
At springs L/20 to L/25
At crown L/40 to L/65
Depth of the deck related to arch span length L L/80 to L/120 a)

a) Deck depth depends on the relative span between vertical supports on the arch.

B2
B2
B2

B1 =
B2
B1 >
B2
B1 ≫
B2

Figure 3.42 Geometry of the arches in the spring sections as the span increases.

intermediate supports usually varies between 30 and 55 m. This spacing depends


very much on the construction system envisaged. If, as usual, the arch is built first
and then the deck, the maximum relative span of the deck depends on the limits of
the system used for its construction (usually movable scaffolding) and on the forces
generated in the different phases of the construction process. The ratio of span to
depth of the deck in relation to these internal columns is usually in the order of
L/14 – L/18 while the relative slenderness of the arc span to the depth of the deck is
in the order of L/80 to L/100, increasing the slenderness as the arch span increases.
As with all HSRBs, one of the fundamental aspects to consider, and also for arch
bridges, is the transmission of the horizontal loads due to the braking forces. One
of the simplest and most common ways to solve the transmission of this load is by
linking the deck and the arch (Figure 3.43).
In this way, the braking forces applied to the deck are transmitted from that con-
nection point to foundations through the arch. It is common for the construction of
arches to be cantilevered with the aid of a temporary cable stay system.
3.5 Structural Types 85

SECTION A–A
A

Figure 3.43 Example of deck to arch connection in the crown.

Figure 3.44 Example of tied arches.

3.5.3.2 Tied Arch Bridges


When the level of the track is very low compared with the ground, it is not possible
to use the type of arches as explained in Section 3.5.3.1. An alternative is to build
the arch above the deck, in a configuration called ‘tied-arch’ or ‘bow-string’. In this
case, the compression of the arch is balanced by the deck (Figure 3.44).
It is usual in this type of bridges that the bridge is formed by two lateral arches that
support the deck. It is also common for the deck to be formed by two longitudinal
side steel beams that are rigidly connected to the arches in the abutment area. The
system is completed with metal crossbeams separated from each other by 1.00 to
3.00 m and an upper slab that directly receives the load from the railway platform.
The arches are usually made in steel and the deck is also usually made of steel for the
longitudinal behaviour. In Table 3.8 a summary of the main dimensions of this type
of bridges is presented.
The usual form of construction of this type of bridges is either by means of tempo-
rary propping, or by building the complete bridge in an area close to the worksite for
its complete transfer by means of a self-propelled modular transporter or by means
of cranes to its final position.

3.5.4 Cable-Supported Bridges


3.5.4.1 Extradosed Bridges
Extradosed bridges are not common in railway bridges; nevertheless, this type
of structures has some qualities that could make them fully appropriate for HSR
Bridges. Those structures are normally built in concrete and have a good stiffness
86 3 Conceptual Design of High-Speed Railway Bridges

Table 3.8 Summary of dimensions for steel tied-arch.

Span range 50 to 150 m


Span/raise 1/5 to 1/7
Depth of the arch related to span length L L/50 to L/60
Depth of the deck related to span length L L/20 to L/4
Concrete deck average equivalent thickness m3 /m2 0.40
Average steel rebars weight in deck per concrete volume kg/m3 250
2
Deck average steel weight kg/m 950 to 1200
Standard self-weight (kN/m) 270 to 300
Standard mean dead load (ballasted track) (kN/m) 170
Standard Permanent Load with ballasted track (kN/m) 440 to 470

Figure 3.45 Example of extradosed bridge.

Table 3.9 Summary of dimensions for extradosed


concrete bridges.

Span range 70 to 150 m


Span/tower height above the deck 1/7 to 1/10
Total depth related to span length L
• Constant depth L/20 to L/25
• Variable depth
– At mid span L/40 to L/35
– At supports L/19 to L/21

that makes them appropriate to HSR Lines. Furthermore, the stays have a small
variation of the stress due to the live load which entails a good behaviour against
fatigue (Figure 3.45).
In Table 3.9 a summary of the main dimensions of this type of bridges is presented.
All the references included come from the Japanese High-Speed Lines.
Extradosed Bridges have a major application in the case of multi-span structures
where they exhibit a substantially higher stiffness compared with the traditional
Cable-Stayed Bridges. More extradosed bridges for HSRL are likely to be built in
the coming years.

3.5.4.2 Cable-Stayed Bridges


In the case of longest spans, a feasible solution could be cable-stayed bridges with
a stiff deck. A way of having the required stiffness with a limited mass is designing
3.5 Structural Types 87

120
120

Figure 3.46 Example of double-level truss deck. ADF – FHECOR – IDEAM.

the deck as a truss steel system. With that kind of solution, the vertical deformation
could be controlled.
For this reason, the decks have to have a significant stiffness. One way of provid-
ing sufficient rigidity without excessively increasing the weight of the structure is to
make a steel truss in which the tracks are located inside the structure.
This solution also allows for the incorporation of traffic on the upper level, thus
generating a two-level structure (Figure 3.46).
As mentioned before, the key problem for HSL Bridges is the relative rotation
at the expansion joints. A way of tackling that problem is by introducing an extra,
shorter lateral span. With that the rotation at the expansion joint will be drastically
reduced (see Figure 3.47).
In Figure 3.48, the proposal for the crossing of the HSL over the Tagus River in
Lisbon is a good example of the importance of the proper shaping of the elements
of such structures. Here, the use of continuous approach viaducts minimises the
rotation problems at the expansion joints. The double-level deck gives the required
stiffness of the system. The minimisation of the masses decreases the seismic loads
and contributions to the general slenderness of the proposal in spite of their colossal
dimensions.

r1

r2

r2 ≪ r1

Figure 3.47 Control of the relative rotation at the expansion joint by introducing an
additional lateral span.
88 3 Conceptual Design of High-Speed Railway Bridges

Figure 3.48 Tagus River Crossing proposal, 550 m main span, DJV
ADF – FHECOR – IDEAM.

17.50

Figure 3.49 HSL Wufengshan Yangtze River Bridge, deck’s cross section.

3.5.4.3 Suspension Bridges


Suspension bridges have a similar problem to cable-stayed bridges, but increased
since in general a cable-stayed system is more rigid than a suspension system. Sus-
pension bridges such as cable-stayed bridges need a very rigid truss deck to meet the
requirements of high-speed lines.
The Wufengshan Bridge in China will be the longest railway bridge in the world
(1092 m). The deck is a double-level truss for road and rail traffic (Figure 3.49) and
has short approach spans to solve the problem of relative rotations at the bridge’s
expansion joints (Figure 3.50).

Figure 3.50 HSL Wufengshan Yangtze River Bridge, hybrid bridges.


3.6 Structural Elements – Substructure 89

Figure 3.51 Cross section


tubular-framed HSRB
(Source: FHECOR).

Figure 3.52 Tubular-framed Ulla Bridge (Source: FHECOR).

3.5.4.4 Hybrid Bridges


Combinations of structural mechanisms to span long obstacles are always a possi-
bility. Tubular structures in combination with arches or frames open a wide range of
new structural systems that could solve the increasing demand of long, strong, and
elegant structures (Figure 3.51). The tubular option provides sufficient stiffness for
high spans where it is competitive with suspended or guyed solutions. It is a feasi-
ble solution when there is no need for a double deck on the deck, i.e. when it is a
rail-only bridge.
In Figure 3.52, a combination of a frame and a tubular deck for a long span viaduct
is shown.

3.6 Structural Elements – Substructure


Here, specialities in substructure (abutments, piers, and fixed points) regarding the
HSR requirements are presented. The design of abutments is highly conditioned by
the presence or not of an expansion joint in the rails, as well as whether the abutment
is the longitudinal fixed point of the deck or not. Likewise, the design of a fixed point
90 3 Conceptual Design of High-Speed Railway Bridges

in some intermediate area of the deck is a specific feature of long high-speed viaducts
and is therefore analysed with special attention in this subchapter. Although not so
specific to HSR bridges, the usual types and configurations of the piers and bear-
ing devices common to this type of bridges are described. Finally, a description and
analysis of the special anti-seismic devices is included.

3.6.1 Abutments
There are basically three types of abutments on HSR bridges.
– With an expansion joint in the deck but without a track expansion joint.
– With an expansion joint in the deck and with a track expansion joint.
– Fixed: without a structure expansion joint.
In all cases, the abutments are designed to be closed to ensure the stability
of the approach wedge. Likewise, except in the case of clear rock, the abutments
for HSR bridges are usually built on piles to avoid significant vertical movements
in the deck – abutment transition. The most significant aspects of each of the
above-mentioned abutment types for HSR bridges are reviewed below.

3.6.1.1 Abutments with Expansion Joint in Structure Only


This is the case of conventional abutments which, in addition to the pressure
of the ground, are subject to the friction associated with the abutment bearings
(Figure 3.53).

3.6.1.2 Abutments with Expansion Joint in Structure and Track


If there is an expansion joint in the track, it is advisable to have an area at the back of
the abutment to house the part of the track expansion joint located in the abutment,
in order to avoid vertical movements in the switches of the joint. The dimensions of
this area depend on the length of the track expansion device (Figure 3.54).
As it is an abutment with a structural expansion joint, the loads transmitted by the
deck correspond to the longitudinal friction of the abutment’s bearing devices and
the transversal forces transmitted through them (Figures 3.55 and 3.56).

3.00 17.00
Ballast
Continuous rail Sub-ballast
0.35
0.40

0.35

Shape layer
0.30
0.60

Structural
expansion 1 1
joint 1 1
Waterproofing
sheet

Cement Gravel Embankment


treated
gravel

Figure 3.53 Longitudinal configuration of an abutment with structural expansion joint


and without track expansion joint.
3.6 Structural Elements – Substructure 91

Ballast
Sub-ballast

0.30
Expansion rail joint 3.00 17.00 Shape layer

0.60
Waterproofing
sheet
1
1
Cement Gravel Embankment
treated
gravel
Drain
tube

Figure 3.54 Longitudinal configuration of an abutment with track expansion joint.

1.95 1.60 15.40 1.30 0.75 15.40 1.60 1.95 1.35


1.30 1.30 1.30
Pot type Pot type
1.97

bearing bearing
1.11

9.36
9.32
7.39

0.50

0.50
0.10

0.10
2.50

2.50

2.50

2.50
8.10 12.75 2.80 2.80 12.75 8.10
0.10
0.10

0.10
0.10

0.10
L pile

Ø1.80 Ø1.80 Ø1.80 Ø1.80 Ø1.80


23.65 23.65
Right elevation Left elevation

VAR. 1.30 15.40 1.30


Space for track expansion joint

VAR. 1.10
VAR

Pot type
bearing
VAR
7.37

0.50
2.50

1.90 12.75 2.80

Ø1.80 Ø1.80
0.10
L pile

1.80 4.50 2.75 6.60 6.60 1.40

Longitudinal section

Figure 3.55 Lateral elevation and centreline cross section of an abutment with track
expansion joint.

3.6.1.3 Fixed Abutments


These abutments are the most complex to be found on railway bridges, as they have
to transmit very significant loads from the deck to the foundation (Figure 3.57).
The materialisation of the fixed point can be achieved by three different types of
connections:

– Fixing the deck to the abutment by using POT-type supports, one fixed and the
other transversally guided. This solution is feasible up to horizontal design loads
of 3000–3500 kN. In these cases, the bearings must be taken from the catalogue as
a function of the vertical design load, and the horizontal forces to be resisted must
be specified in the drawings (Figure 3.58).
92 3 Conceptual Design of High-Speed Railway Bridges

14.00

0.30
2% 2%

3.40

10.00
4.80
0.50
8.72

4.25
12.02

5.25
1.15 1.15

13.80
1.50
26.50
3.00

4.50
(MIN.)

Ø1.50 Ø1.50 Ø1.50 Ø1.50


0.10

1.80
16.50
4.50

1.80
1.50 4.50 4.50 4.50 1.50
16.50
Elevation

0.60
4.50
1.50
1.50 4.50 4.50 4.50 1.50
16.50
Plan

Figure 3.56 Front elevation and plan view of an abutment with track expansion joint.

Ballast Ballast
Sub-ballast
Continuous
Shape layer 21.00 3.00
0 .3 0

rail
0. 60

Waterproofing
1 Sheet
2 1
2
Drain tube
Embankment Gravel Cement
treated
gravel

Figure 3.57 Longitudinal configuration of a fixed abutment.

– Fixing the deck to the abutment by using brackets that are lowered under the deck
and that bump against the abutment (Figure 3.59).
– Fixing the deck to the abutment by means of bars or prestressed tendons. This
type is usually used in viaducts where the final fixed point is located at interme-
diate points of the viaduct and it is necessary to fix the deck to the abutments
temporarily during construction (Figure 3.60). (In Spain, the Railway Authority
ADIF prefers the use of prestressed tendons for permanent fixing).
Other aspects to take into account in the design of those elements are as follows:
If the fixed point is realised with POT bearings, there must be sufficient horizontal
3.6 Structural Elements – Substructure 93

Pots

Pot
transversally Fixed pot
guided

Pot
Fixed pot
transversally
guided

Figure 3.58 Detail of longitudinal fixing with POT bearings.

Pots

Neoprene

Neoprene

Pot Pot free


longitudinally
guided

Neoprene

Pot Pot free


longitudinally
guided

Figure 3.59 Detail of longitudinal fixing with brackets.


94 3 Conceptual Design of High-Speed Railway Bridges

Prestressed rebars
Prestressed rebars

Neoprene

Neoprene

Pots
Pot
longitudinally Pot free
guided

Neoprene

Pot Pot free


longitudinally
guided Prestressed
rebars

Figure 3.60 Detail of longitudinal fixing with prestress.

reinforcement under these bearings capable of introducing the horizontal force into
the body of the abutments.
If the fixed point is materialised with brackets, the space designed to house the
longitudinal stopper must allow the replacement of the neoprene that is placed in
the bracket to ensure proper contact between the bracket and the body of the abut-
ment. One possible method is to place the neoprene on a plate that is screwed to the
abutment, so that the screws are accessible from the sides of the bracket.
When designing the deck span next to the fixed abutment, the moment at the
end due to the eccentricity of the horizontal load must be taken into account. This
moment is attenuated until it becomes practically nil in the second pier.
Since the horizontal loads to be transmitted to the ground are very important,
it may be interesting to use a friction slab (see Figure 3.61). In the case of abut-
ments with direct foundations, it may be necessary to use a friction slab to avoid
over-dimensioning the footing. If the foundation is deep (most common), a friction
slab is usually used so as not to penalise the piles which, if not available, should
be able to transmit all the horizontal force to the ground. For the verification of
stability against sliding, it is usually considered that the friction slab rubs only on
its lower face. The combined operation of slab and piles is complex. A common
way to define the length of the friction slab required is to design it to resist all the
horizontal force in SLS, considering only the combined effect of the slab and the
piles for fulfilling the ULS.
For maintenance details of abutments see Chapter 7.
3.6 Structural Elements – Substructure 95

VAR.
5.40 5.40
1.30 1.80 1.80 1.30 1.80 1.80

VAR. (3.01–3.15)

3.40
VAR.

2.60

12.16
5.75
3.15
6.00 5.40 5.10 6.00 5.40 5.10

2.00
2.00

1.00

3.00
3.00
0.10 100

(MIN.)

(MIN.)

(MIN.)
0.10

0.10

0.10
(MIN.)

1.50 4.50 4.50 4.50 1.50 1.50 4.50 4.50 4.50 1.50
10.00 16.50 10.00 16.50

Seccion longitudinal Seccion longitudinal

13.80
0.50
3.00

5.70

6.00
12.02
5.48

3.01

11.40 5.10
2.00
1.00

3.00
(MIN.)

(MIN.)
0.10

0.10

1.50 4.50 4.50 4.50 1.50


10.00 16.50

Elevation

Figure 3.61 Fixed abutments with friction slab.

3.6.2 Piers
Due to the magnitude of the vertical loads, railway bridge piers are often very robust.
In general, the piers have to transmit a significant longitudinal force, which corre-
sponds to the longitudinal deck force produced by the friction of the support devices
located at their head. Transversally, the governing load is usually the transverse wind
or the earthquake in the case of bridges located in areas with seismic activity.
Except when no fractured rock appears on the surface, in which case the founda-
tion of the piers is normally by footings, the foundation of piers of HSR bridges is
pile-caps with large-diameter piles.
In the case of a concrete slab or prefabricated deck, which corresponds to the
low span range, the piers are usually solid, as they rarely reach 20–25 m in height
(Figure 3.62).
For larger spans, associated with higher piers, those elements are usually hollow,
with constant cross sections up to 25–30 m in height and with variable dimen-
sions for higher values. See details for inspection and maintenance in Chapter 7
(Figure 3.63).

3.6.3 Bearings
Except for very special cases, the support devices for HSR’s bridge decks are of the
‘POT’ type made of neoprene confined in a metal cup and with a Teflon finish so that
the friction transmitted to the infrastructure is as small as possible (Figure 3.64).
96 3 Conceptual Design of High-Speed Railway Bridges

A A
5.20 1.90
Pot type Pot type
bearing bearing
Hf

Hf
3.40 5.20 3.40 3.25 1.90 3.25
2.70

2.70
0.10

0.10
Ø1.80 Ø1.80 Ø1.80 Ø1.80 Ø1.80

1.50 4.50 4.50 1.50 1.50 5.40 1.50


12.00 8.40

5.20
1.56 1.04 1.04 1.56

Space for
0.10 jacking

0.10
12.00

1.40
1.50 4.50 4.50 1.50

0.10
1.50

0.10
Pot type
.80 .8 0 .8 0 bearing
Ø1 Ø1 Ø1 1.90 1.90
2.70

1.90 1.90
8.40
1.90

1.40

5.20
2.70

2.60 2.60
Pot type bearing
5.20
.80 .80 .80
Ø1 Ø1 Ø1
1.50

1.90

1.40
0.25

0.20

Figure 3.62 Example of solid pier for a concrete slab deck type.

When smaller size is required, spherical bearings are a good alternative


(Figure 3.65).
The service life of these elements is less than expected for the whole bridge, so it is
necessary to provide for their replacement. The deck, piers, and abutments must be
designed for this operation. See Chapter 7 for details for maintenance and bearing
replacement.

3.6.3.1 General Bearing Layout


In this case of simply supported bridges, each section of deck is fixed longitudinally
to one of the piers on which it rests and is free longitudinally in the other support.
3.6 Structural Elements – Substructure 97

3.51 8.53

Bearing Inspection hole Bearing Inspection hole

4.75

4.75
1.20

1.20
1 1 1 1
50 50
17,5 17,5
H=32.484

H=32.484

4.81 12.24
4.00

4.00
0.10

0.10

7.00 7.00 8.00 8.00


14.00 16.00

Longitudinal elevation Transverse elevation

16.00
8.00 8.00

Bearing
0.60

12.24
0.40

7.00

1.20

0.40
14.00
VAR.

0.60
4.81

3.51

0.60
0.40
7.00
0.40

8.53
0.60

Inspection
VAR. hole
Cross section
Plant

Figure 3.63 Example of hollow pier with variable dimensions in height.

In both piers, the deck must be transversely constrained, so the support arrangement
is as shown in Figure 3.66.
In the case of redundant decks, the support in piers is only transversal, so one of
the bearings must be guided longitudinally and the other free. In the case of short
bridges without earthquakes, where the fixed point on the abutment can be mate-
rialised by ‘pot’ supports, one of them will be fixed and the other will be guided
transversally (see Figure 3.67). If the longitudinal loads to be transmitted to the fixed
point on the abutment are large, the support configuration will be: one free and one
longitudinally guided support.
98 3 Conceptual Design of High-Speed Railway Bridges

Leveling
B Leveling
B Slide plate
mortar
mortar Slide plate
Slide plate Slide plate
Guide

Leveling
Leveling Mortar
A mortar A A
SECTION B-B SECTION B-B

GUIDE
Leveling Leveling
Slide plate Slide plate
Mortar mortar

B
B Leveling Leveling
SECTION A-A Mortar
Plant Plant SECTION A-A mortar

Generally mobile multidirectional pot bearing Unilateraly mobile unidirectional pot bearing

Leveling
Mortar

A A
H

Leveling
Mortar

PLANT SECTION A-A

Fixed pot bearing

Figure 3.64 Example of type of POT bearings.

Leveling Leveling
Slide plate
mortar mortar

Leveling Leveling
mortar mortar
SECTION A-A SECTION A-A

Slide plate

A A A A

Plant Plant

Fixed pot spherical bearing Generally mobile pot spherical bearing

Figure 3.65 Example of spherical bearing.

Structural Structural Structural


expansion joint expansion joint expansion joint

Fixed Longitudinally Fixed Longitudinally Fixed Longitudinally


guided guided guided

Transversally Transversally Transversally


Free Free Free
guided guided guided

Figure 3.66 Bearing arrangement simply supported decks.


3.7 Seismic Design 99

Expansion joint

Longitudinally Longitudinally Longitudinally Longitudinally


Fixed
guided guided guided guided

Tranversally Free Free Free Free


guided pot

Longitudinal "Movable abutmen"


fixed point
"fixed abutmen"

Longitudinal Expansion joint


fixed point

Longitudinally Longitudinally Longitudinally Longitudinally


Fixed
guided guided guided guided

Longitudinal loads Tranversally Free Free Free Free


transmission guided pot
prestresed rebars
"Fixed abutmen" "Movable abutmen"

Figure 3.67 Bearing arrangement continuous decks.

3.7 Seismic Design

3.7.1 Seismic Design Strategies


Seismic design strategies for any type of bridges are based on three main tools:

– Decrease the mass of the bridge, to decrease the seismic action.


– Increase the flexibility of the structure, whenever possible, to increase the
vibration period and consequently decrease the seismic action. To do this, the
flexibility must be such that the resulting period has to exceed the plateau of
the response spectrum.
– Dissipate seismic energy by increasing the damping of the structural system.

In principle, not all of these strategies can be used in HSRB.


One of the possible strategies to use is the reduction of the self-weight of the deck.
And therefore, in zones of high seismicity it is often more convenient to use steel or
composite (steel and concrete) decks because of their lower self-weight compared
to concrete decks. It is also possible to use energy-dissipating devices as will be
explained later. However, it is very difficult to reduce the rigidity of the infrastructure
as a minimum longitudinal rigidity must be guaranteed to ensure rail traffic safety.
Sections 3.7.2–3.7.4 explain in more detail the practical application of these concepts
depending on the bridge articulation.

3.7.2 Seismic Behaviour and Deck Articulation


There are two fundamental deck configurations as shown in Section 3.6.3.1. On
the one hand there are the simply supported solutions and on the other hand the
continuous decks. In the first case, each pier will have to resist the seismic force
corresponding to one span of the deck both longitudinally and transversally. If the
deck is continuous and with a longitudinal fixed point, it will have to resist the lon-
gitudinal seismic force of the whole deck, while transversely each pier will have to
100 3 Conceptual Design of High-Speed Railway Bridges

resist the seismic load corresponding to a single span. Therefore, it is the longitu-
dinal behaviour that is fundamentally different in each of the situations as will be
explained below.

3.7.3 Longitudinal Behaviour


3.7.3.1 Simply Supported Spans
In this case, each span is linked longitudinally to a pier. Therefore, each pier will
have to resist the longitudinal seismic action of only one span. In general, the
longitudinal connection between the deck and the pier will be fixed, either by
means of longitudinally aligned support devices or by means of a longitudinal
stopper.
If the piers do not have the resilience to withstand the seismic load, longitudi-
nal dampers could be used in each support to dissipate some of the seismic energy,
making the system more flexible to reduce the force on the substructure.

3.7.3.2 Continuous Deck


In the case of continuous decks, all the seismic force associated with the deck
mass is concentrated in the longitudinal fixed point. The resulting load is very
important which makes the design of the fixed points practically unfeasible in the
case of medium and long viaducts. For this reason, the most common alternative
in seismic areas is to have a system of longitudinal dampers that allow the max-
imum value of the force transmitted by the deck to the fixed point to be limited
(Figure 3.68).
The design of the dampers must be such that there is no relative movement
between the deck and the fixed point for the braking and starting situation. There-
fore, the load produced by the movement of the damper system must be greater
than this force with a certain safety coefficient (usually 1.20).
One of the problems with the use of dampers at the fixed point is that the safety
of rail traffic is compromised when it starts to move relative to the deck and fixed
point. For this reason, the dimensioning force of the dampers, and therefore of the
fixed point, must be greater than the force produced by frequent earthquakes, and

Adjustable
expansion joint

Deck Abutment

2 Pot bearings
Cement
2 Viscous treated
fluid dampers gravel

Abutment

Figure 3.68 Hydraulic dampers between deck and abutment.


3.7 Seismic Design 101

only in the case of ultimate or collapse earthquakes does the damper system come
into operation by reducing the seismic load on the fixed point.

3.7.4 Transversal Behaviour


3.7.4.1 Introduction
In the event of a transverse earthquake, each element of the infrastructure (piers
or abutments) will have to resist a transverse load corresponding approximately
to the force generated by the earthquake in the section of the deck that rests
on them.
However, the response to the earthquake will be very different in the case of simply
supported decks than in the case of continuous decks.
In the first case, as there are structural expansion joints between each span, rela-
tive rotations can occur between them and therefore the track, which is continuous,
has to be able to follow the relative movements between decks [12]. Rail traffic can
therefore be affected by this effect (Figure 3.69).
In the case of continuous viaducts, this effect does not occur, since the continuity
of the deck ensures the compatibility of the vertical axis turns along the deck. Only
at the points where there is a structural expansion joint, generally in abutments, can
there be a relative rotation of the vertical axis.
In this type of bridges, as the deck is continuous, the force that is transmitted to
each pile depends not only on the spans that rest on the pile and therefore on the
relative mass of that section of deck, but also on the relative stiffness of each pile
in relation to the stiffness of the whole and on its position more or less close to the
points that do not move.
Both in the case of simply supported decks and in the case of continuous decks,
there are two fundamental ways of transmitting the load to the infrastructure: either
by means of transversal stoppers or by means of a system of dampers. These two
possibilities are described below.

3.7.4.2 Fixed Transversal Support


It should be noted that the horizontal loads that can be transmitted by a standard
POT-type device in the direction in which the guide materialises are of the order
of 30% of the maximum vertical load of a bearing. Nevertheless, in special cases,

Earthquake

Openin g
Pounding g Openin Pounding

g Pounding Pounding Openin


Openin g

Figure 3.69 Relative vertical rotation axis during transversal seismic.


102 3 Conceptual Design of High-Speed Railway Bridges

100% of the vertical load can be reached. This means that in the case of bridges
with significant transversal loads (i.e. seismic loads), transversal stoppers are
necessary to transmit these loads, as the bearing devices do not have sufficient
capacity.
Figure 3.70 shows an example of a central transversal stopper in the case of a
viaduct with a deck formed by two u-beams
In any case, in structures in a seismic area it is essential to have these transversal
stoppers in order to avoid the exit of the deck support in its support in piers.

3.7.4.3 Transversal Damping Systems


In the event that it is necessary to lower the transverse load in the piers due to the
earthquake, it is possible to use a damping system (Figure 3.71).
As in the longitudinal case, the damping system only works in the ultimate
earthquake situation. Both for the normal operation of the bridge and for frequent
earthquakes, the transverse forces must be resisted by the guides of the support

Neoprene

Transversal
Stoppers

Figure 3.70 Transversal stoppers at piers.

Steel connection

Transversally
Free bearing

Transversal damper

Figure 3.71 Transversal damping system in piers.


3.7 Seismic Design 103

Figure 3.72 Special transverse reinforced bearing device with guaranteed maximum
horizontal force seismic devices.

devices. In these cases it is very important to design these elements or side stoppers
that work as a fuse so that once the threshold of the design force defined by the
designer is exceeded, the system is released from these elements and the transverse
dampers start working (Figure 3.72).
There are a variety of devices for damping bridges. In general, there are two main
groups of devices: those in which the damping is done in the same bearings and
those that are independent elements of the vertical support of the bridge.
The first include devices such as High Dumper Rubber Bearings, Lead Rubber
Bearings, Hysteretic damping with steel damping devices, and Pendulum devices.
The second type is the viscous hydraulic dampers devices.
Section 3.7.4.4 explain the practical application of these types of systems to HSRB
Damping through the bearings.
Neither High Dumper Rubber Bearings nor Lead Rubber Bearings are usually used
in HSRBs as these bearings have a shorter service life than POT-type bearings. Steel
damping and pendulum devices could be used but in combination with transversal
stoppers (Figures 3.73 and 3.74).
As explained above, transversal stoppers can be designed in two ways. Either
they are completely rigid, in which case there is no transversal damping, or they
function as a fuse that allows, in this case, the damper-supporting device to move
laterally therefore dissipating seismic energy and modifying the lateral flexibility of
the system.
104 3 Conceptual Design of High-Speed Railway Bridges

Dynamic longitudinal movement Dynamic transversal movement

Viaduct longitudinal axe Viaduct longitudinal axe

FD
FY

SY SD

Figure 3.73 Example of steel damping bearings.

2500

2000

1500
e=D
1000
Keff
500
V V 0
–800

–600

–400

–200

200

400

600

800
0

–500

–1000

–1500

–2000

–2500

Figure 3.74 Pendulum devices.

3.7.4.4 Damping Devices Plus Bearings


The other design alternative consists of separating the normal bridge support,
which is materialised in POT-type support bearings, from the damping system for
horizontal seismic loads. In this case, this type of viscous hydraulic dampers devices
is usually used to achieve the system’s damping (Figure 3.75).
As mentioned above, if the system is damped transversely, it is necessary to have
a transversal restraint system for the service earthquakes and for the usual service
situation of the bridge.
3.7 Seismic Design 105

Figure 3.75 Example of hydraulic viscodamper.

From a longitudinal point of view, the damper devices are usually placed in one
of the abutments. In this case, it is necessary to have an elastic system to achieve the
re-centring of the deck in the post-seismic situation. If it is not used, the deck could
be displaced from its position before the earthquake.
One of the usual ways of solving the problem in long viaducts is to link the deck to
one or two central piers by means of fixed supports so that the piers act as an elastic
re-centring element. If the central piers are significantly higher than the rest, it is
usual to connect two piers to the deck. On the other hand, if the supports are rigid,
only one of them can be connected. In this case, it is necessary to check that the
flexibility of the pier is such that it does not absorb an excessive longitudinal force
that makes the solution unviable (Figure 3.76).
If this solution were not feasible, it would be necessary to have some neoprene
in addition to the general system of vertical bearings that would allow the elastic
re-centring of the system (Figure 3.77).

A1 P-1 P-2 P-3 P-4 P-5 A2


Expansion
joint

Bearings Bearings Bearings Fixed Bearings Bearings

Bearings
Hydraulic
dampers

Figure 3.76 Example of re-centring with flexible – fixed piles in the centre of the bridge.

Figure 3.77 Example with re-centring by additional neoprene (plan and lateral view).
106 3 Conceptual Design of High-Speed Railway Bridges

3.8 Worked Example


3.8.1 Introduction: Aim and Data
This section develops an example of the conceptual design of a high-speed viaduct.
The idea is to define the dimensions of a viaduct, about 400 m long, which is needed
to cross a U-shaped valley at a height above the valley floor of about 30 m.
Sections 3.8.1.1–3.8.1.5 include the data required for the preliminary design of the
viaduct.

3.8.1.1 Topography
The attached Figure 3.78 defines the plan and longitudinal profile of the route in the
area where a viaduct is planned to be built.

3.8.1.2 Plan and Elevation


The structure is located in a section in which the plan corresponds to a circular align-
ment of 7500 m radius, and a longitudinal profile with a constant slope of 0.43%.

3.8.1.3 Railway Platform Section – Project Speed


The viaduct will be designed for a double-track railway platform with a design speed
of 300 km/h (Figure 3.79).

Elevation

P.C. +640.00 528 + 700 528 + 800 528 + 900 529 + 000

E-1
P.K. 528 + 631

528 + 700 528 + 800 528 + 900 529 + 000


528 + 600

Figure 3.78 Plan and elevation.

14.00
6.25 6.25
5.70 5.70
5.05 5.05

2.35 2.35
0.78

3
0.14

h 3
2
0.82

2
2% 2%
0.40

0.56

Figure 3.79 Railway platform.


3.8 Worked Example 107

3.8.1.4 Water Flood Level


The viaduct saves the permanent watercourse of a river and its floodplain. The
maximum level of the 500-year return period flooding reaches 670 m.

3.8.1.5 Preliminary Geotechnical Data


The foundations located on the slopes can be directed at 0.6 N/mm2 in the rocky
substrate which is considered to be located at a depth of 4.00 m below the natural
ground level.
The foundations in the river floodplain will be made by large-diameter piles
working at an equivalent permissible stress of 6 N/mm2 .

3.8.2 Methodology
A logical sequence to determine the position of the bridge on the ground and define
its spans and dimensions of the main elements could have the following steps:
– Analysis of existing information
– Preliminary location of the abutments: definition of the total length of the viaduct
and position of the longitudinal fixed point
– Distribution of spans
– Deck pre-dimensioning
– Bearings
– Pre-dimensioning of pier elevations and foundations

3.8.3 Critical Analysis of Existing Information


When dealing with the situation of the structure in the landscape, it is necessary to
have some basic information listed below, the review and assimilation of which is
essential in the initial phase of the design process:
– Situation plan at least on a 1/1000 scale and if possible, a topographic survey at
1/500 of the critical zones,
– Plan layout and elevation
– Section of the railway platform in the structure
– Speed of the line project
– Critical elements for the fitting with indication of the clearances to be respected:
– Roads
– Railways
– Channels or rivers
– Constructions
– Utilities
– Specific requirements included in the Environmental Impact Statement
– General and particular landscape
– Preliminary geotechnical report:
– Foundation conditions
– Excavation slopes
108 3 Conceptual Design of High-Speed Railway Bridges

This information constitutes the minimum data necessary to carry out the prelim-
inary design of the structure.
Later, for the development of the detailed design, this information will have to
be completed in some aspects such as the specific conditions of each particular
foundation of the structure.

3.8.4 Determination of the Length of the Viaduct, Selection of the


Fixed Point
It is usual to start with the locating of the abutments, which are the meeting point of
the railway with the ground. From a first trial, an approximate value of the length of
the viaduct will be obtained and consequently, the need or not to have track expan-
sion joints and the characteristics and position of the longitudinal fixed point of the
bridge.
The abutments should have a moderate height, if possible, of less than 10 or 12 m
(15 m as maximum for instance for the Spanish Railway Authority ADIF). Normally,
one of the abutments is the fixed point of the structure. In that case, the abutment
to be chosen, as a fixed point of the bridge, should be the lowest and with the best
geotechnical conditions.
In order to have abutments of 10 m or less in height, abutment number 2 must be
placed around the PK 529+000 (Figure 3.80), and abutment number 1 close to the
PK 528+630. Therefore, the structure has a length of approximately 370 m.
As the viaduct is more than 100 m long, a rail joint will be required. For this rea-
son, one of the abutments must be fixed and the opposite must have the required
geometry to allocate the track expansion device.
In this case, since both the heights and the geotechnical conditions are similar in
both abutments, the fixed point could be made indistinctly at either end of the deck.

529 + 000

Figure 3.80 Abutment n∘ 2.


3.8 Worked Example 109

3.8.5 Span Distribution


Knowing, approximately, the total length of the bridge, the distribution of spans
will be made, with a constant sequence, except in extreme spans, provided that the
position of the obstacles to be overcome allows it.
An ideal value for the ratio of side span to central spans is 0.80–0.85. Taking into
account this ratio and the total length of the structure, different distributions of the
spans may be made, all of which meet the above criteria.
The relationship between span/piles height, the constructive possibilities (means,
accesses, etc.) will serve for the selection of the typology of the deck, within the
possibilities analysed previously.
In the present case, the presence of a permanent watercourse means that it is
almost mandatory to create a centred span with this river, which should be around
50–70 m wide to provide the necessary protection when the foundations of the piers
adjacent to the watercourse are built.
Among the possible span distribution, the one with a span of 65 m has been cho-
sen, since it has the minimum number of piles in the flood zone, an area where the
foundations are much more expensive as they need pile foundations.
The 65 m span is located at the upper limit of the construction by self-launching
formwork. Likewise, this solution allows it to be pushed, given the characteristics
of the alignment (in a curve with a constant radius and with a constant longitudinal
slope), and as the length of the viaduct is sufficient, longer than 300 m, to make
the construction of the thrust park necessary for the manufacture of the deck
economically feasible.
Thus, a 370 m long solution has been chosen, consisting of 6 spans, with 55.00 + 4
times 65.00 + 55.00 spans between support axes (Figures 3.81 and 3.82).

3.8.6 Deck Pre-dimensioning


Based on the span and the construction system foreseen, the typology of the deck
can be defined, as well as its deck and the geometry of the cross section.

368.00
E-1 P-1 P-2 P-3 P-4 P-5 E-2
P.K. 528 + 631 55.00 P.K. 528 + 685 65.00 P.K. 528 + 750 65.00 P.K. 528 + 815 65.00 P.K. 528 + 880 65.00 P.K. 528 + 945 55.00 P.K. 528 + 999

Cota rasante Aparato de dilatación


Z = 694.867 0.43 % DE VÍA:±145 mm

RIO

P.C. + 640.00 528 + 700 528 + 800 528 + 900 529 + 000

Figure 3.81 Viaduct elevation.

368.00

P-1 P-2 P-3 P-4 P-5 E-2


E-1
P.K. 528+631 54.00 P.K. 528+685 65.00 P.K. 528+750 65.00 P.K. 528+815 65.00 P.K. 528+880 65.00 P.K. 528+945 54.00 P.K. 528+999
7.00
14.00
7.00

Figure 3.82 Plan view of the structure.


110 3 Conceptual Design of High-Speed Railway Bridges

As the main span is 65 m, the two most feasible alternatives are a prestressed con-
crete deck with a box section or a double steel beam with a concrete slab deck.
Of the two solutions, the pre-stressed box solution was chosen because, as the
height of the piles is moderate, the friction during a possible push does not produce
excessive longitudinal moments in the pier. Furthermore, as the structure is not in a
seismic zone, the advantage of the decrease in mass associated with the steel solution
is not a great advantage.
For this reason, a prestressed concrete section with a deck of 4.45 m has been
planned, which means a span-to-deck ratio of L/15, which is very suitable for con-
struction using the push technique (Figure 3.83).
The standard cross section of the deck chosen is the typical one for a
post-tensioned concrete box with a constant deck. The bottom width of the
box is 6.00 m, with sloping side webs and 3.00 m long end cantilevers.
The box section is made up of a lower slab 0.36 m thick in the centre of the span,
which increases to 0.75 m in the area on piers with gussets at the junction with the
cores. The upper slab is of variable thickness due to the transverse slope with a min-
imum of 0.35 m in the centre of the slab. This slab is limited in its union with the
webs. The thickness of the webs is 0.50 m, which allows for parallel pre-stressing
cable pairs (Figures 3.84 and 3.85).
The estimated average dead weight of the deck is 297.5 kN/m, and the dead loads
to be considered are those included in Table 3.10.

14.00
6.25 6.25
5.70 5.70
5.05 5.05

2.35 2.35
0.78

3 3
0.14

2
0.20
0.82

2
2% 2%
0.40

0.56
0.40

0.21

0.13

0.58 0.58
3.60 1.22 3.20 1.22 3.60
3.70

4.36

0.55
0.35

5.70
1.12 3.00 1.12
0.40
0.62

6.50

Figure 3.83 Cross section of the deck.


3.8 Worked Example 111

14.00

0.14
0.20

0.40
0.13

0.21

0.21
1.80 3.20 1.80

0.55 0.55
3.41

4.50
4.36
0.71 3.20 0.71

0.35
0.62

0.40
3.60 0.55 5.70 0.55 3.60

Figure 3.84 Deck cross section in the centre of span.

14.00

0.14
0.20

1.00
0.13

2.98
3.41

4.50
4.36
2.75

1.80
1.80
2.10
0.75
0.62

3.60 5.70 3.60


0.55 0.55

Figure 3.85 Deck cross sections in piers supports.

Table 3.10 Dead loads.

Element Load (kN/m)

Ballast guard walls 2 × 0.20 × 0.50 × 25 = 5.0


Imposts 2 × 0.24 × 25 = 12.0
Gutter systems 2 × 3 = 6.0
Railway sleepers and rails (for the 2 × (4.8/0.60 + 2 × 0.6) = 18.4
whole of the two tracks)
Handrails 2 × 0.35 = 0.7

Ballast
Max 10 × 0.60 × 20 × 1.3 = 156.0
Med 10 × 0.60 × 20 = 120.0
Min 10 × 0.60 × 20 × 0.70 = 84.0

Total
Max 198.1
Med 162.1
Min 126.1
112 3 Conceptual Design of High-Speed Railway Bridges

3.8.7 Pre-design of the Infrastructure


3.8.7.1 Fixed Point
In relation to the performance against horizontal loads, a bridge has been designed
and anchored lengthwise in one of the abutments (Abutment n∘ 2). The fixing of the
deck to the abutment is made by means of pre-stressed bars.

3.8.7.2 Bearings
The transversal loads are going to be resisted in all the piers and abutments, and the
torsional forces are also resisted in each support (piers or abutments). Therefore, in
each pier or abutment two bearings are used.
One is unidirectional, longitudinally guided POT, to transfer the transversal
forces and the other is a multidirectional POT-type device, free in both directions,
longitudinal and transversal (Figures 3.86 and 3.87).

3.8.7.3 Abutments
The abutments are of the closed type. Due to the length of this viaduct, a track
expansion device has been planned for Abutment number 1 (Figure 3.88).
In Abutment number 2 the fixed point of the deck is materialised by connecting
the deck to the abutment wall by pre-stressed bars. A friction slab has been installed
to prevent the sliding of the abutment (Figure 3.89).

Pot free
Longitudinally guided pot

E-1 P-1 P-2 P-3 P-4 P-5 E-2

Figure 3.86 General bearings layout.

D1 D2
ZS
D1 D2 ZNMB
ZS
A C
E.sup.
ZNMB
E.sup.
C

GUÍA
C

E.inf.

E.inf. ZAP.
ØA
D1

ZAP. 50 50
Section C–C
D1

ØA
50 50
B B Section A–A
ØA D ØA
D
D

E
ZS GUÍA E
ZNMB ZS
E.sup. ZNMB
E.sup.
D2

D2
C

E.inf.
E.inf.
ZAP.
ØA ZAP.
ØA
50 Section B–B 50
50 50
Section D–D
A C
Plan Plan
Longitudinally guided pot Free pot bearing

Figure 3.87 Longitudinally guided POT and multidirectional free POT bearing.
3.8 Worked Example 113

1.95 1.60 15.40 1.30


1.30
Pot type
1.97

bearing

1.11

9.32
7.39

0.50
2.50

2.50
8.10 0.10 12.75 2.80
0.10

0.10
L pile

Ø1.80 Ø1.80
23.65

Figure 3.88 Abutment n∘ 1.

Ballast Ballast
Sub-ballast
Shape layer Continuous rail
0.30

21.00 3.00
0.60

Waterproofing
1 sheet
2 1
2
Drain tube

Embankment Gravel Cement


treated
gravel

Figure 3.89 Abutment n∘ 2 (Fixed point).

3.8.7.4 Piers
The piers on which the deck rests are made up of a single shaft in a box section with
external dimensions of 6.00 m wide and a variable deck, the minimum value of this
being 2.40 m at the crown and increasing with height at a rate of 4 cm/m. The wall
thickness is 0.40 m. At the top, the section has a solid zone of 1.50 m. The height of
these piers is between 17 and 24 m.
The foundations of piers 1 and 5 are direct by footings, while piers 2, 3, and 4
have deep foundations, consisting of a pile cap of 6 piles each that rest on rock
(Figures 3.90 and 3.91).
114 3 Conceptual Design of High-Speed Railway Bridges

A
6.50 A
Pot free 2.00 2.50 2.00 2.40

1.30
1.30

1.50
1.50
Longitudinally
guided pot 6.50

0.20
0.20

0.03 1.00 0.03 0.10 0.10


0.20 6.10 0.20

0.20
0.40

0.05
0.05
0.05 0.05

0.40
Variable
Variable

0.40
0.05

0.05
0.05 0.05

0.40 0.40 0.40

0.20
0.40 0.40
B B B B
Section B-B
Scale 1: 50
H

H
6.50
0.20 6.10 0.20

0.20
2.55 1.00 2.55

2.40
2.00
0.20
2.20 2.20

3.25 4.00 3.25 2.00 2.50 2.00


0.25 0.25
3.00 6.50 3.00
View “A”
Scale 1: 50
1.00

1.00
2.50

2.50
1.50

1.50
(Min.)
0.10
(Min.)

6.50 6.50 5.25 5.25


0.10

13.00 10.50

Figure 3.90 Piers 1 and 5.

6.50
2.00 2.50 2.00 2.40
1.30

1.30
1.50

1.50
0.20

0.20

0.03 0.03 0.10 0.10


1.00

6.50
0.20

0.20 6.10 0.20

0.40
0.05

0.05

0.05 0.05
VARIABLE

VARIABLE

0.40

0.40

0.40 0.40 0.05


0.05

0.05

0.05

0.40
0.20
H

SECCIÓN B-B
50

6.50
0.20 6.10 0.20
0.20

2.55 1.00 2.55


2.40
2.00

3.25 6.50 3.25 VARIABLE VARIABLE VARIABLE

0.40
0.20

2.20 2.20

2.00 2.50 2.00

VISTA POR A
3.00

3.00

50
(MIN.)

(MIN.)
0.10

0.10

1.50 5.00 5.00 1.50 1.50 5.00 1.50


13.00 8.00
ALZADO FRONTAL. ALZADO LATERAL.
75 75

Figure 3.91 Piers 2 to 4.


References 115

References

1 Manterola J., Astiz M.A., and Martínez A. (1999). Puentes de Ferrocarril de Alta
Velocidad Revista de Obras Públicas n∘ 3386 Madrid.
2 ADIF. Norma ADIF Plataforma NAP 2-0-0.1. Puentes y viaductos ferroviarios. 2a
Edición 2019.
3 Romo, J., Corres, H., and Pérez, A. (2005). High speed railway bridges in seismic
areas. In: IABSE Symposium: Structures and Extreme Events, Lisbon, Portugal.
4 Rui-Wamba, J. (2020). Teoría Unificada de las Estructuras y Cimientos. Barcelona:
Editorial Reverté.
5 Romo J. (2010). Riudellots High Speed Line Bridge IABSE Symposium: Large
Structures and Infrastructures for Environmentally Constrained and Urbanised
Areas, Venice, Italy.
6 Leonhardt, F. (1983). Bridges Aesthetic and Design. Deutsche Verlags-Anstalt
GmbH.
7 Zeller, W. and Saul, R. (1991). Long span bridges of the new railroad lines in
Germany IABSE Reports.
8 Schlaich, J., Schmitt, V., Marx, S. et al. (2008). Leitfaden Gestalten von
Eisenbahnbrücken, 1e. Berlin: DB Netze AG.
9 Schlaich, M. (2012). Integral Railway Bridges in Germany. 22nd Dresdener
Brückenbausymposium. Dresden.
10 Martin J.L. (1999). Viaduc sur l’A7 à Bonpas. Ponts métalliques Bulletin 19
OTUA.
11 Plu B., Durot F., and Teisseire J. (1999). Viaduc de la Toulubre. Ponts métalliques
Bulletin 19 OTUA.
12 Priestly, M.J.N., Seible, F., and Calvi, G.M. (1996). Seismic Design and Retrofit of
Bridges. John Wiley & Sons Inc.
117

Design Basis
José Romo

4.1 Introduction
This chapter sets out the specific actions for rail bridges in general and for bridges
for high-speed lines in particular. It also indicates the criteria for combining these
actions for the various checks to be carried out.
In general, the different design regulations for bridges are based on a series of
general criteria and on the verification of the structure according to the theory of
Limit States: Service and Ultimate. The former (S.L.S.) are those which, if exceeded,
will cause the structure to cease to fulfil the purpose for which it was designed
(whether for functional, personal comfort, or aesthetic reasons), without this lead-
ing to its collapse, while the latter (U.L.S.) are those which, if exceeded, will lead to
the exhaustion or collapse of the structure or part of it.
In addition, and specifically for railway bridges, special checks must be carried
out to ensure the safety of the railway traffic running on the bridge. These checks
are described later in this chapter.
Likewise, and specifically for accidental load situations, the design of the struc-
ture must be robust enough so that damage from these accidental actions, such as
those corresponding to a derailment, does not cause disproportionate damage to the
bridge.

4.2 Design Situations


A design situation in the project of a structure is usually defined as the situations that
characterise a certain period of time during which the factors affecting its security
can be considered unchanged.
In general and in accordance with the usual standards, these are project situations:
– Persistent situations: They correspond to the normal conditions of use of the
structure during its useful life.

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
118 4 Design Basis

– Transitional situations: They occur during the construction, inspection, or


conservation of the structure. Their duration can be considered to be one year.
– Accidental situations: These correspond to exceptional conditions of stress on
the structure, e.g. fires, explosions, impacts, or derailments. They can be consid-
ered to have an instantaneous duration.
– Seismic situations: These correspond to exceptional conditions of stress on the
structure due to earthquakes.
In general, rail bridges are intended to have a lifespan of 100 years, which is the
period of time from their entry into service, during which they should fulfil the func-
tion for which they were built, with adequate conservation, but without requiring
rehabilitation operations. And therefore the safety format, including the charac-
teristic value of the actions, their combination, and enhancement coefficients, is
theoretically calibrated in order to achieve that service life.

4.3 Rail Traffic Actions and Other Actions Specific


of Railway Bridges

This chapter includes the specific loads and actions of railway bridges in general and
bridges for high-speed railway lines.
Firstly, permanent actions are defined which include, in addition to the own
weights of the elements, some special loading situations such as the partial removal
of the ballast in an area of the bridge and the actions due to overhead line equipment
and other railway infrastructure and equipment.
Secondly, actions due to railway operations are explained, including:
– vertical loads,
– dynamic effects,
– centrifugal forces,
– nosing forces,
– traction and braking forces,
– aerodynamic actions from passing trains.
Finally, accidental actions as the effect of traffic derailment are presented.

4.3.1 Permanent Loads


4.3.1.1 Self-Weight
The load is deduced from the theoretical geometry of the structure and considered
for the densities of the different structural materials. The usual values for the
different types of deck are given in Chapter 2.

4.3.1.2 Dead Loads


These are due to the non-resistant elements. These loads include:
– Ballast: The density to consider of the ballast is normally 20 kN/m3 . To take
into account the variability of its placement, a nominal value is usually adopted
4.3 Rail Traffic Actions and Other Actions Specific of Railway Bridges 119

(the one indicated in the drawings of the project) and a lower and higher value
have to be considered. Thus:
– Nominal value, determined by the theoretical thicknesses defined in the
project. In general, in the absence of specific data, a thickness of 0.50 m can be
considered for UIC track widths.
– Lower value (Gk,inf), obtained by decreasing the nominal value by thirty percent
(30%).
– Higher value (Gk,sup ), obtained by increasing the nominal value by thirty per-
cent (30%).
– Rails: frequently of the UIC-60 type: According to Table A.6 of EN 1991-1-1 [1], a
load of 1.2 kN/m must be considered and per track.
– Prestressed concrete sleepers: According to Table A.6 of EN 1991-1-1, a load of
4.8 kN/m per track is to be considered.
– Ballast walls: Load obtained from their geometry.
– Sidewalks: Load obtained from its geometry and density of materials.
– Piping gutters: According to experience a load of 3.0 kN/m and track should be
considered.
– Catenary: Normally the load is transmitted pointwise coincident with the cate-
nary poles. It is not a critical load for the bridge and its value is of the order of
0.2–0.3 kN/m for a two-track bridge.
– Railings: Load obtained from the geometry and density of the materials. In the
case of a steel-type railing, a load of 0.35 kN/m per side can be considered
(Figure 4.1).

4.3.1.3 Partial Ballast Removal


The structure should be designed for the forces that can occur as a result of main-
tenance on one track while trains are running on another track. Checks should be
carried out at U.L.S and S.L.U.
Some codes define for instance two situations to be considered as follows:
– Half of the ballast thickness is removed from a track on all or part of the bridge to
produce the most unfavourable effect (Figure 4.2).
– The entire ballast thickness is removed in a track over a length of 15.0 m and in
the area of the bridge that produces the most unfavourable effect (Figure 4.3).

4.3.2 Variable Loads


4.3.2.1 Vertical Live Loads
Rail traffic actions are defined in the various codes. In general, the following types
of charges are defined
Figure 4.1 Typical cross
section.
120 4 Design Basis

100% Figure 4.2 Partial ballast


50% removal (Situation 1).

100% Figure 4.3 Partial ballast


0% in 15.00 m removal (Situation 2).

250 kN 250 kN 250 kN 250 kN


q = 80 kN/m qvk = 80 kN/m
vk

0.80 1.60 1.60 1.60 0.80

Figure 4.4 Example of loads representing normal traffic on main lines (source EN
1991-2) [2].

qvk q
vk

a c a

Load qvk a c
model (kN/m) (m) (m)
SW / 0 133 15,0 5,3
SW / 2 150 25,0 7,0

Figure 4.5 Example of heavy loads models (source EN 1991-2) [2].

– Loads representing normal traffic on the main lines. For example the load model
71 in EN 1991-2 (Figure 4.4).
– Heavy load models: As is the case with the SW/2 in EN 1991-2 for instance
(Figure 4.5).
Passenger train loads at speeds exceeding 200 km/h, such as HSML freight models.
These are the load models used in dynamic analyses.
– Load model ‘train without load’ representing the effects of an unloaded train.
When combined with the action of wind, for example.

Depending on the line loads, the corresponding load values are usually amplified
or reduced (e.g. in EN 1991-2 loads 71 and SW/2) with a classification coefficient.
4.3 Rail Traffic Actions and Other Actions Specific of Railway Bridges 121

Qvi
qv1 + qv2 (1)
Qv1 + Qv2

qv1QV1 qv2QV2
Qvi/2

4:1
e Qvi/4 Qvi/4

(2)

r a a b

Figure 4.6 Transverse eccentricity, load distribution in sleepers, and load diffusion through
the ballast (source EN 1991-2).

The codes indicate the eccentricities to take into account possible transverse load
imbalances between rails. For example in EN 1991-2 an eccentricity e = r/18 is
indicated, where r is the distance between rails which in the case of 1435 mm results
in 80 mm. The standards also indicate how the point loads are distributed between
the individual sleepers, as well as the distribution of the loads on the ballast. For
example, in EN 1991-2, 50% of the load is considered for the sleeper underneath
and 25% for the adjacent sleepers. In this standard, the load distribution through
the ballast is considered to be at a 4 : 1 gradient (see Figure 4.6)

4.3.2.2 Traction and Braking Forces


The starting and braking forces are very important on the bridges of high-speed lines
as they often condition the dimensioning of the infrastructure (piers and abutments).
The braking and starting actions of railway vehicles are assimilated to horizontal
forces, parallel to the track, uniformly distributed along a certain length and applied
at the level of the median running plane.
For example, in EN 1991-2, the overall value of these actions, for a track, is as
high as 6000 kN for the braking force and 1000 kN for the starting force on normal
tracks. These high load values, combined with the limitations of the relative horizon-
tal displacements (see Chapter 1), mean that bridges on high-speed lines require a
fixed longitudinal point, either an abutment or an intermediate pier (see Chapter 3)
(Figure 4.7).
It is important to note that when the track is continuous at one or both ends of the
bridge, only a part of the starting or braking force is transmitted through the deck to

Rail Deck Ballast / Rail expansion


track effect joint

Figure 4.7 Braking load transmission through a continuous rail.


122 4 Design Basis

the supports; the rest of the force is transmitted through the track and resisted behind
the abutments. The fraction of the force transmitted through the deck to the supports
should be determined taking into account the interaction of track–structure.
Braking and starting forces not only condition the bridge substructure but also
increase the stresses on the rails. To determine the value of these stresses it will be
necessary to carry out an analysis of the track–structure interaction as described in
detail in Chapter 6.

4.3.2.3 Centrifugal Forces


For bridges with curved track, the action of the centrifugal force caused by the move-
ment of railway vehicles along the track shall be assimilated to a set of horizontal
forces, either punctual or uniformly distributed, perpendicular to the axis of the
track and applied at a certain height above the median running plane. In EN 1991-2
it is indicated, for example, that this force shall be considered at a typical height
of 1.80 m. For some types of traffic, for example with double height of containers, a
higher value should be considered for the height of application of the load. The value
of the centrifugal force is proportional to the square of the train speed, inversely pro-
portional to the radius of curvature of the track on the ground, and proportional to
the mass of the train.
In bridges with two tracks, the centrifugal force can act on one or both tracks, and
therefore one or the other situation must be considered according to whether it is
more unfavourable for the element and effect under study.
The characteristic value of the centrifugal force is
v2
Qtk = ⋅ (f ⋅ Qvk ) (4.1)
gR
with

Qtk Characteristic values of the centrifugal force corresponding to the vertical loads
(kN, kN/m).
Qvk Values of vertical loads, not affected by any impact coefficient, in (kN, kN/m).
v Maximum speed (m/s).
R Radius of curvature (m).
f Reduction factor

The reduction value f takes into account the influence length Lf of curved track
on the deck which is most unfavourable for the design of the element under consid-
eration, as well as the speed of the train (see EN-1991-2).
For example, when designing a bridge pier with continuous spans of 40 m,
Lf = 40 m, the value of the reduction factor f for a speed of 350 km/h is 0.40. If the
radius is 8000 m (standard value for a speed line of 350 km/h) the centrifugal force is:
(350∕3.6)2
Qtk = ⋅ (0.40 ⋅ Qvk ) = 0.048 Qvk (4.2)
g ⋅ 8000
Therefore the centrifugal force is close to the 5% of the vertical live load acting on
the pier.
4.3 Rail Traffic Actions and Other Actions Specific of Railway Bridges 123

4.3.2.4 Nosing Forces


This is another typical action of railway bridges. It is a single horizontal and punctual
force perpendicular to the axis of the track, applied at the top of the rail and towards
the outside of it and must always be combined with the vertical traffic load. Its value
is normally under about 100 kN and therefore does not usually condition the general
design, although it can influence the local design of some secondary element.

4.3.2.5 Aerodynamic Actions from Passing Trains


Railway traffic generates alternative pressure and suction waves that travel with the
train and that occur mainly at its head and tail. They affect the elements close to the
track, both structural (top webs and wings in trough or box sections; uprights and
diagonals in lattice-type decks; piers or cables in cable-stayed or suspended bridges,
etc.) and non-structural (anti-noise screens, barriers, canopies, etc.).
The magnitude of these actions depends on the square of the train speed, its aero-
dynamics, and the shape and position of the loaded elements. These actions are
assimilated to static pressure and suction, and are used to check the ultimate limit
states of breakage and fatigue.
The above actions are included in EN 1991-2 and reach transverse pressure val-
ues of 1.8 kN/m2 when a train passes at 300 km/h and at a distance of 2.30 m or
2.5 kN/m2 of vertical pressure or suction when there is a horizontal element located
5 m above the track level.
These values are amplified by 2 in the first few meters closest to the beginning or
end of these elements (Figure 4.8).

4.3.2.6 Thermal Actions


In railway bridges the effect of the track–structure interaction has to be considered
to determine the stresses in the structure due to thermal effects. Beside its effect on
the structure thermal actions also generate stress on the rails, which must be taken
into account when checking the rails (see Chapter 1).

Surface of
structure
q
1k

Surface of
structure
5.00

±q1k
≤ 5.00

5.00

a
g
q1k

Figure 4.8 Pressure distribution on a side web in a U-shaped section.


124 4 Design Basis

4.3.2.7 Bearing Friction


The forces due to the friction of the supports bearings condition the longitudinal
behaviour of the structure.
The values of the friction coefficients in the support bearing are very important for
studying both the interaction between the track and the structure, the distribution
of the braking loads and the seismic behaviour of the bridge, and in general for the
longitudinal behaviour of the bridge.
In general the supports will be longitudinally sliding and a value of the forces of
extreme friction must be considered.
It is usual to consider extreme values (favourable and unfavourable) to cover
different design situations safely. In addition, it is common to take into account
the number of support devices when considering these values of extreme friction
coefficients, as will be seen later.
For instance, EN 1337-1 [3] indicated the coefficients of friction values 𝜇 a and 𝜇 r
as follows:
𝜇a = 0.5𝜇max (1 + 𝛼)
𝜇r = 0.5𝜇max (1 − 𝛼)
where 𝜇 a is the adverse coefficient of friction and 𝜇 r is the favourable coefficient of
friction.
For sliding supports with PTFE, a maximum friction coefficient 𝜇max of 3% is
usually considered, regardless of the pressure to which the support is subjected. 𝛼
is a factor that depends on the partial number of supports (n) that exert either a
favourable or an unfavourable constraint on the element under study.
The value of 𝛼 to be used according to the standard is 1 for 4 or less supports, 0.5 for
more than 10 supports, and an interpolated value between 1 and 0.5 for intermediate
cases. This factor will generally adopt different values depending on whether 𝜇a or
𝜇 r is being calculated.

4.3.3 Dynamics Effects


4.3.3.1 Introduction
The static forces induced in a bridge are increased and decreased under the effects
of mobile traffic, for several reasons. Firstly, the rapid loading due to the speed of the
traffic running on the structure and the inertial response of the structure (impact);
secondly, the passage of successive loads with a similar cadence that can excite the
structure and, in certain circumstances, create resonance (when the excitation fre-
quency or a multiple of it coincides with the natural frequency of the structure or a
multiple of it, there is the possibility that the vibrations caused by the successive axes
passing through the structure are excessive); and in addition the variations in wheel
loads caused by track or vehicle imperfections (including wheel irregularities).
The principal factors which influence dynamic behaviour are: the speed of traffic,
the span length of the element, the mass of the structure, the natural frequencies
and mode shapes along the line of the track, the number of axles, axle loads and the
spacing of the axles, the damping of the system, the vertical irregularities in the track,
4.3 Rail Traffic Actions and Other Actions Specific of Railway Bridges 125

the mass and suspension characteristics, the presence of regularly spaced supports
of the deck slab or track, the vehicle imperfections, and the dynamic characteristics
of the track.

4.3.3.2 Consideration of Dynamic Effects


The dynamic response of the structure as a result of railway loads may be close to the
resonance values. This implies that the real stresses may increase significantly with
respect to the static values, and furthermore that the induced vertical acceleration
may create railway safety problems and/or loss of comfort for the users.
On railway bridges on high-speed lines, dynamic analysis is required. The purpose
of that analysis is two-fold. On the one hand it is aimed at identifying the velocity
at which resonant frequencies occur and determine a dynamic load factor to be
applied to the real trains in order to assess deflections and forces on the structure.
The structure is to be designed for the worst case of the critical train with its impact
factor which can now go over the value of 2.00 due to resonance, or the design
equivalent train increased by the envelope dynamic factor. On the other hand
dynamic analysis is carried out to determine the maximum accelerations which
are used to assess problems of track instability, deconsolidation of ballast, and
passenger comfort. Chapter 5 is entirely devoted to these aspects.

4.3.4 Railway Vehicle Derailment


Railway bridges must be designed so that, in the event of derailment, there is no
overturning or overall collapse of the structure. For example, in standard EN 1991-2,
two accidental situations are considered. The first assumes that the derailed vehicle
is left with one of its alignments between the rails and the second assumes that
the railway vehicle is supported by a single alignment of wheels on the edge of the
railway platform (Figure 4.9).

4.3.4.1 Railway Vehicle Impacts


Impact Against Structural Elements of the Deck The load-bearing elements of the deck
on the sides or above the deck must be able to withstand the impact of a derailed
railway vehicle. The impact is usually assimilated to a horizontal static force parallel
to the track, or to a perpendicular force to the track, and with a direction towards the

(1) (1)

(2) (2)
α × 0,7 × LM 71 α × 0,7 × LM 71 α × 0,7 × LM 71

(1)

(3) (2) (3) (2) 0.45

Figure 4.9 Accidental derailment situations.


126 4 Design Basis

Figure 4.10 Impact in the


upper structure.

1.80 TYP

outside of the deck, applied at a height of usually 1.80 m above the running surface
of the rail closest to the element.
These loads can be critical in the case of non-massive elements, such as cables,
diagonals, and other truss elements.
The values of the longitudinal loads to be considered depend on the different
codes, reaching 10 MN in the longitudinal direction and 3.5 MN in the transversal
direction if the distance between the element and the axis of the nearest track is less
than 3 m. These values are zero when the element is located more than 5 m from the
axis of the nearest track.
The importance of these loads means that in general, if the bridge has sensitive
elements, the design must either move these elements away from the tracks (more
than 5 m), or use containment measures or increase the resistance of the element
itself of the structure.
Although these loads are accidental they are usually combined with permanent
loads and live loads. In particular, it is usually established that after the impact and
if the damaged element is an element of a certain strength, the bridge must be able to
withstand the permanent actions and static live loads, considering that the impacted
element has lost half of its resistance capacity. In the case of less robust elements,
such as cables, it is usually required that the bridge is able to withstand the per-
manent actions and static use live loads without counting on the resistance of the
element on which it has been impacted (Figure 4.10).

Impact on Bridge Substructure Elements On railway bridges over railway tracks, the
piers faces and other deck supports located less than 5.0 m from the nearest track
axis should support forces of similar values as indicated in Section 4.3.4.1.

4.4 Application of Traffic Loads on Railway Bridges


4.4.1 General
To determine the most unfavourable effect of traffic actions, different scenarios are
established depending on the type of loads considered and the check to be carried
out. In general, the codes establish between different load situations, some when it
comes to carrying out the classic structural checks and others for checking relating
to the safety of rail traffic and checks on passenger comfort.
4.4 Application of Traffic Loads on Railway Bridges 127

4.4.1.1 Load Situations for Structural Design


For example, the EN-1990 indicates how to calculate the worst situations in bridges
when it comes to evaluating the effect of live loads type 71, SW0, and SW2 in bridges
with 2 or more tracks which are indicated in detail in Section 4.4.2.

4.4.1.2 Load Situations for Limit State and Associated Acceptance Criteria
In the case of checks relating to rail traffic safety and passenger comfort, the codes
usually establish the number of loaded tracks to be considered in each case.
In general, when it comes to checking the safety of rail traffic, all loaded tracks
are considered, while for the comfort of users, only one track is usually considered
loaded.
Thus, for example in the Eurocode EN 1990 [4], the traffic safety checks include:
deck twist, vertical and horizontal deformation of the deck checks is carried out
considering the worst situation even those where all tracks are loaded even if there
are three or more.
In the same code, for the verification of the combined response of the structure and
track to variable actions including limits to vertical and longitudinal displacement
of the end of a deck a maximum of two loaded tracks are considered. On the other
hand, when checking the vertical acceleration of the deck and the comfort of the
passengers, this standard considers only one loaded track.

4.4.2 Groups of Loads


4.4.2.1 Characteristic Values of Multicomponent Action
As seen previously, rail traffic involves the application of vertical loads (F V ), as well
as horizontal loads in the longitudinal direction (F HL ) and in the transverse direction
to the bridge axis (F HT ).
It is clear that the probability that the maximum of all forces due to traffic will
occur simultaneously is very low. That is why the standards establish load groups,
which are mutually exclusive, that usually are considered as a single variable char-
acteristic action for combination with non-traffic loads.
In the case of the Eurocode EN 1991-2 the vertical loads FV correspond to the
LM71, SW/0, HSLM, SW/2, or the unloaded train loads; the longitudinal forces (F HL )
to the traction and braking forces, and the transverse forces (F HT ) to the centrifugal
and nosing forces.
In that code the single variable action described above is obtained adding the
forces as shown as indicated in Table 4.1.
Table 4.1 Assessment of groups of loads (Characteristic values of
the multicomponents actions).

Assessment of group loads for rail traffic FV FHL FHT

Vertical with max. longitudinal 1.0 1.0 0.5


Vertical with max. transversal 1.0 0.5 1.0
Additional load case 0.75 0.75 0.75
128 4 Design Basis

Other representative values of the multicomponent actions.


For the various checks, the loads due to the rail traffic must be combined with
other actions. The combination factors depend, in general, on the type of vertical live
load considered. Thus, for example, the Eurocode EN 1990 [4] establishes a value of
𝜓 0 of 0.80 for the LM71 and SW/0 loads, a value of 𝜓 0 of 1.00 for the HSML and the
unloaded train, and a zero value for SW/2.

4.5 Traffic Loads for Fatigue


The different standards establish the reference overloads to carry out these checks
which will generally try to reflect the characteristics of the traffic planned for the line.
In order to carry out the fatigue checks, the dynamic load amplification factors must
be taken into account, whether they are real trains or equivalent loads. Basics for
fatigue assessment of railway bridges are defined in Eurocode EN 1991-2 Annex D,
for example [5].

4.6 Verifications Regarding Deformation and Vibrations


for Railway Bridges
Checks on bridge deformations shall be performed for traffic safety purposes for the
following items:
– Vertical accelerations of the deck to avoid ballast instability and unacceptable
reduction in wheel rail contact forces.
– Vertical deflection of the deck throughout each span to ensure acceptable vertical
track radii and generally robust structures.
– Unrestrained uplift at the bearings to avoid premature bearing failure vertical
deflection of the end of the deck beyond bearings to avoid destabilising the track,
limit uplift forces on rail fastening systems, and limit additional rail stresses.
– Twist of the deck measured along the centre line of each track on the approaches
to a bridge and across a bridge to minimise the risk of train derailment.
– Rotation of the ends of each deck about a transverse axis or the relative total rota-
tion between adjacent deck ends to limit additional rail stresses.
– Limit uplift forces on rail fastening systems and limit angular discontinuity at
expansion devices and switch blades.
– Longitudinal displacement of the end of the upper surface of the deck due to
longitudinal displacement and rotation of the deck end to limit additional rail
stresses and minimise disturbance to track ballast and adjacent track formation.
– Horizontal transverse deflection to ensure acceptable horizontal track radii.
– Horizontal rotation of a deck about a vertical axis at ends of a deck to ensure
acceptable horizontal track geometry and passenger comfort.
– Limits on the first natural frequency of lateral vibration of the span to avoid
the occurrence of resonance between the lateral motion of vehicles on their
suspension and the bridge.
4.7 Worked Example 129

In addition to these checks relating to the safety of rail traffic, other aspects must
be considered, such as those relating to passenger comfort, which are specified by
establishing a limit for vertical accelerations and vertical deformations of the deck.
These special checks related to traffic safety and passenger comfort are included
in Chapters 2 and 5.

4.7 Worked Example


4.7.1 Introduction
The first step is to establish the main actions for the design of the viaduct studied in
Chapter 3, which is represented in Figures 4.11 and 4.12.
Later, the bearings, the fixed abutment, and the piers of the viaduct are preliminary
dimensioned. The methodology used consists of the following stages.
Loads Eurocode 1 was used to obtain both the vertical loads (permanent loads,
serviceability overloads, and impact coefficient) and the horizontal loads (braking
and acceleration forces, and wind action).

368.00
E-1 55.00 P-1 65.00 P-2 65.00 P-3 65.00 P-4 65.00 P-5 55.00 E-2

RIO

Figure 4.11 Elevation.

14.00
6.25 6.25

Catenary 2.35 2.35 Catenary


axis axis
0.14
0.20
0.40

0.13
0.21

0.58 3.20 0.58


3.60 1.22 1.22 3.60
3.70

0.55
0.35

5.70
1.12 3.00 1.12
0.40
0.62

6.50

Figure 4.12 Cross section.


130 4 Design Basis

The actions were then combined according to Eurocode 0 (basis of structural


design) and Eurocode 1 part 6 (rail traffic actions and other actions specifically for
railway bridges).

4.7.1.1 Calculation of Reactions at Bearings: Pre-dimensioning


Based on the actions and the structural configuration, approximate values for the
reactions were obtained, which were used for the preliminary dimensioning of the
various deck bearings.

4.7.1.2 Calculation of Forces and Preliminary Design of the Fixed Abutment


As mentioned above, one of the special design elements of a high-speed railway
bridge is the point at which the deck has to be longitudinally fixed to the infrastruc-
ture to avoid longitudinal relative movement between deck and piers and abutments.
In this example, the fixed point is at one of the abutments, which will therefore
be subjected to a significant longitudinal load. This section sets out the loads to be
considered and the preliminary design of the abutment.

4.7.2 Actions
4.7.2.1 Vertical Loads
Permanent Loads The permanent loads are, on the one hand, the own weight of the
deck and, on the other hand, the weight of the track and other elements of the railway
superstructure.
For the self-weight a typical cross section area of 12.22 m2 is used; therefore
SW = 12.22 ⋅ 25 kN/m3 = 305.6 kN/m.
For the deck loads (ballast, track, etc.) the maximum and minimum values
considered are: DLmax = 198.1 kN/m and DLmin = 126.1 kN/m.

Live Loads According to the EN 1991-2 for lines with no heavy trains load types 71
and SW/0 have to be considered in the design of the bridge (Figures 4.13 and 4.14).

250 kN 250 kN 250 kN 250 kN


qvk = 80 kN/m q = 80 kN/m
vk

0.80 1.60 1.60 1.60 0.80

Figure 4.13 Load type 71.

qvk qvk

a c a

Load qvk a c
model [kN/m] [m] [m]
SW / 0 133 15,0 5,3

Figure 4.14 Load type SW/0.


4.7 Worked Example 131

Due to the length of the span a single track with load type 71 has a total bending
moment of M = 50 150 kN/m and a maximum shear of V = 3060 kN whereas for the
load type SW/0 the total bending moment is M = 44 650 and the shear V = 2900 kN.
Therefore the load type 71 is the governing live load.
For the design of the bridge a classification factor of 1.21 has been consid-
ered as it is a line with slightly heavier traffic than a normal line. Therefore the
LL1 = 1.21 ⋅ 80 ⋅ 2 = 193.6 kN/m.
For the sidewalks a live load of 5 kN/m2 has been considered. In combination with
the railway load this live load in the sidewalks has a combination factor of 0.50;
therefore, the load to be considered is: LL2 = 2 ⋅ 1.80 ⋅ 5 kN/m2 ⋅ 0.50 = 9.0 kN/m.
For the purpose of preliminary dimensioning, the locomotive weight included in
the 71 type load has been simplified by an equivalent point load of 1180 kN which is
used in conjunction with the linear load of 80 kN/m.

QLL = 1.21 ⋅ (1000–80 ⋅ 6.40) ⋅ 2 = 1180 kN

Therefore the equivalent load is a uniform load, in the whole span, of 202.5 kN/m
and a point load of 1180 kN.

Dynamic Factor 𝚽 To take into account the dynamic amplification of the loads the
impact coefficient has been considered according to EN 1991-2.
The equivalent LΦ is given by the expression: LΦ = k ⋅ Lm = 1.50 ⋅ (54 + 4 ⋅
65 + 54)/6 = 92 m.
The value of the amplification factor Φ2 assuming a carefully maintained the track
is according to EN 1991-2 (6.4.5).
Φ2 = √ 1.44 + 0.82 = 0.973, and therefore the value to be considered is Φ2 = 1.00:
92−0.2
no dynamic amplification for the resistance of the structural members has to be con-
sidered.

Live Load Eccentricity Eccentricity of the vertical loads LM71 and SW/0 is calculated
according to EN 1991-2 (6.3.5) (Figure 4.15).
1.45
e = 2.35 + = 2.43 m
18

4.7.2.2 Horizontal Forces


Actions Due to Traction and Braking According to EN 1991-2 (6.5.3) and considering
the classification factor of 𝛼 = 1.21.

2.35

1.435

Figure 4.15 Load eccentricity.


132 4 Design Basis

The traction force considering La,b = 368 m (total length of the viaduct)
Qlak = 33 kN∕m ⋅ 368 m ≤ 1000 kN => Qlak = 1000 kN
And the braking force:
Qlbk = 20 kN∕m ⋅ 368 m ≤ 6000 kN => Qlbk = 6000 kN
Therefore considering that the train on one track accelerates while on the
opposite track it brakes, the value of the total longitudinal load is equal to:
Qlk = 1.21 ⋅ 7000 = 8470 kN.

Centrifugal Forces The value of the characteristic centrifugal force is obtained as


follows EN 1991-2 (6.5.1):
v2
qtk = (f .qvk ) with v = 300 km/h, r = 7500 m, f = 0.365 according to Table 6.7
127r
3002
qtk = (0.365.qvk ) = 0.034qvk
127.7500
Centrifugal forces consider the classification factor 1.21 as for the live load due to
railway traffic.

4.7.2.3 Wind Speed


Basic Wind Speed Wind load has been analysed according to EN 1991-1-4.
vb,0 = 28 m/s (basic wind velocity for the zone of location of the structure for a
return period of 50 years).
Considering the probability of annual exceedance for 100 years, Cprob = 1.04
vb = 29.12 m∕s

Mean Wind Mean wind is determined using the expression: vm (z) = cr (z). c0 (z) vb
according to EN 1991-4-4 (4.3)
where
c0 (z) is the orography factor: in this case c0 (z) = 1.10 (valley)
cr (z) is the roughness factor
The terrain category of the site is type II.
Therefore kr = 0.19, zo = 0.05 m, zmin = 2 m.
Considering a height z = 25 m
z
cr = kr ⋅ ln = 1.18
z0
Wind turbulence factor
√ √
7kz 7 ⋅ 0.19
cg = 1 + = 1+ = 1.42
cz ct 1.18 ⋅ 1.1

Peak Wind Speed

vbp = 29.12 ⋅ 1.10 ⋅ 1.18 ⋅ 1.42 = 53.6 m∕s


4.7 Worked Example 133

Wind Force in the Deck


Transversal wind
d
Drag coefficient cd = 2.5 − 0.3 with L = 14 m
L
– Without live load
d = 4.44 + 0.60 + 0.20 = 5.24 m (deck height, ballast, and rails)

14
cd = 2.5 − 0.3 ⋅ = 1.70
5.4
– With live load
d = 5.24 + 4.00 = 9.24 m
(deck height, ballast, rails, and UIC train height = 4.00 m)
14
cd = 2.5 − 0.3 ⋅ = 2.05
9.24
Unloaded state (Figure 4.16)
2a Railing:
Table 8.1 EN 1991-1-4 (open parapets on both sides)
A = 4.44 + 0.60 = 5.04 m < 5.24 (deck + ballast height + rail)
( ) ( )
1 2 1
F = cd ⋅ A 𝜌vc = 1.70 ⋅ 5.24 ⋅ 1.25 ⋅ 53.62 ⋅ 10−3
2 2
= 16.0 kN∕m
Loaded state (Figure 4.17)
( )
1
F = 2.05 ⋅ 9.24 ⋅ ⋅ 1.25 ⋅ 53.62 10−3 = 34.01 kN∕m
2

4.7.3 Calculation of Reactions at Bearings: Pre-dimensioning


To obtain a preliminary value of the forces of the bearings a distance of 4.00 m
between axes of the bearings in a pier has been considered.

4.7.3.1 Vertical Forces


In order to determine the distribution of the vertical loads on the supporting struc-
tures, the reactions in each pier due to the vertical loads have been calculated.

Figure 4.16 Deck geometry 14.00


unloaded state.
2.00
4.44
134 4 Design Basis

4.00

9.24
0.80 (0.60 balasto
+0.20 m)

4.44
Figure 4.17 Deck loaded state.

To obtain the load on each bearing due to the centred loads, the corresponding
axial force in the pier was divided by 2.
Nbearing = Npier ∕2
In the case of eccentricity due to vertical loads, the load on each bearing was
obtained as follows:
Nbearing = Npier ∕2 ± Mpier ∕4.00
Being 4.00 the distance between bearings.

4.7.3.2 Centrifugal Forces


The axial force per bearing can be obtained as follows (Figure 4.18):
Nbearing = (1.80 + 0.8 + 4.44) ⋅ Qtk ∕4.00 = 1.76 Qtk

QtK
1.80

0.60 + 0.20
4.44

Nbearing S Nbearing
4.00

Figure 4.18 Centrifugal free action.


4.7 Worked Example 135

If Qv is the vertical load associated with the centrifugal force, and being
Qtk = 0.034 Qvk
The axial force per bearing is N bearing = ±0.034 ⋅ 1.76 = 0.06 Qvk (Tables 4.2 and 4.3).

4.7.3.3 Wind at Unloaded State


( )
5.24 65
Nbearing = ±16.0 ⋅ 10−3 ⋅ ⋅ = ±0.68 MN (central piers)
2 4.00
(65+54)
2
N = ±16.0 ⋅ 10−3 ⋅ 2.62 ⋅ = ±0.62 MN (lateral piers)
4.00
(54)
2
N = ±16.0 ⋅ 10−3 ⋅ 2.62 ⋅ = ±0.28 MN (Abutments)
4.00

4.7.3.4 Wind with Live Load

F = 34.01 kN∕m
9.24
h= = 4.62 m Mτ = L ⋅ 34 ⋅ 10−3 ⋅ 4.62 = 0.157L
2
N = ±2.55 MN (central piers)
N = ±2.34 MN (lateral piers)
N = ±1.06 MN (abutments)

4.7.3.5 Reactions in Pier Heads

Table 4.2 Reaction in pier heads due to permanent loads and transversally symmetrical
vertical live loads.

Reactions (MN) Abutment Pier 1/5 Pier 2/4 Pier 3

Self-weight 6.22 20.25 19.82 19.87


305.6 kN/m
Dead load max. 4.04 13.12 12.85 12.89
198.1 kN/m
Full live load max 5.05 14.62 15.22 15.41
202.6 kN/m + 1.18 MN 1.18 1.18 1.18 1.18
6.23 15.80 16.40 16.59

Table 4.3 Reaction in pier heads due to transversally asymmetrical vertical live loads.

Reactions per abutment/pier per


transversally asymmetric live load Abutment Pier 1/5 Pier 2/4 Pier 3

N (MN) 3.11 7.90 8.20 8.30


M t (MN.m) e = 2.43 m 7.57 19.20 19.93 20.16
136 4 Design Basis

4.7.3.6 Transversal Wind Bearings Reactions


As combined with live load 𝜓 0 = 0.75 corresponding to F wk as F wk > F w ** because
vb0 = 28 m/s > vb ** = 25 m/s. Other non-traffic loads:

Central piers N = ±2.55 MN ⋅ 0.75 = 1.91 MN

Lateral piers N = ±2.34 MN ⋅ 0.75 = 1.76 MN

Abutments N = ±1.06 MN ⋅ 0.75 = 0.80 MN

4.7.3.7 Loads per Bearings


Combining the vertical loads and the effect of the transversal loads the maximum
axial force in the bearings is calculated (Table 4.4).
Therefore, POT bearings of 10 000 kN at the abutments and 30 000 kN at the piers
are provided. The general layout of the bearings is shown in Figure 4.19. In each pier
one of the pots is longitudinally guided to resist the transversal forces and the other
is free. The same disposition is used in the abutments.

Table 4.4 Maximum axial Reaction in pier heads due to transversally asymmetrical
vertical live loads.

Abutment Pier 1/5 Pier 2/4 Pier 3

Permanent load 5.13 16.69 16.34 16.38


LL alternative 1 Full vertical live load 3.11 7.90 8.20 8.30
Centrifugal force (full ±0.37 ±0.70 ±0.85 ±0.86
vertical live load)
LL alternative 2 Asymmetrical live 3.45 8.75 9.08 9.19
loads (max, min) −0.33 −0.85 −0.88 −0.89
Centrifugal force ±0.19 ±0.47 ±0.49 ±0.50
(asymmetric live loads)
Wind (𝜓 = 0.60) ±0.80 ±1.76 ±1.91 ±1.91
Maximum per 9.56 27.67 27.82 27.98
bearing

Pot longitudinally Pot longitudinally Pot longitudinally


guided 10 mn guided 30 mn guided 10 mn
A1 P1 P2 P3 P4 P5 A2

Longitudinal
connection

‘Movable Pot free 30 mn Pot free 10 mn ‘Fixed


abutmen’ Pot free 10 mn
abutmen’

Figure 4.19 Bearings layout.


4.7 Worked Example 137

4.7.4 Fixed Abutment Loads


4.7.4.1 Introduction
As mentioned above, one of the special points of any high-speed railway bridge is the
design of the fixed point. In this case, abutment 2 is the place where the longitudinal
connection of the deck to the bridge infrastructure is made.
The worst situation for the stirrup is when a pull occurs due to horizontal tensile
and braking loads trying to slide the stirrup.
There are several techniques to improve the stability against sliding of the abut-
ment, such as the construction of anti-slip wedges, ground anchors, or friction slabs.
In this example it has been decided to use a friction slab, which is essentially
an extension of the footing on the land side in order to have a higher vertical load
(the weight of the embankment) and thus increase the friction to improve the sta-
bility of the abutment.
It is therefore a question of calculating the length that the friction slab needs
to be in order to comply with the slip safety condition. This starts by determining
the forces acting on the abutment and then defining the required length L of the
friction slab.
The safety criterion against sliding is a safety of 1.50 for the situation without
braking and 1.35 in the case of braking and traction loads.
In Figure 4.20 the cross section of the abutment is shown.

4.7.4.2 Loads Transmitted by the Deck


The loads transmitted from the deck to the abutment are:

– Vertical reaction due to permanent loads of the deck.


– Vertical reaction due to rail traffic.
– Horizontal force due to the friction of the POT bearings in the piers and abutments
(Figure 4.21).
– Horizontal force due to traction and braking.
– Horizontal force due to the variation of temperature between rails and deck
structure.

VA
2.00
HT
c.o.g
4.75

Deck
3.00

1.85
7.90

9.90

HA
0.30

3.15

5.00 5.35 9.65


2.00

Ancho
3.00

15.50
1.00

A
20.00 ¿L?

Figure 4.20 Transversal cross section of the abutment.


138 4 Design Basis

A1 P1 P2 P3 P4 P5 A2

Friction forces
54.00 65.00 65.00 65.00 65.00 54.00

Figure 4.21 Friction forces.

Table 4.5 Forces transmitted by the deck.

Vertical (MN) Horizontal (MN)

Permanent loads 20.37 ⋅ (0.30 + 0.18) = 9.67


Friction forces due to 368 ⋅ (0.30 + 0.18) ⋅ 0.03 = 5.24
permanent loads
Live load (including the 24.91 ⋅ 0.20 + 1.18 = 6.23
locomotives load)
Braking and traction 8.46
forces
Temperature 2.00

These forces are listed in Table 4.5. For the determination of the friction forces of
the POTs a friction coefficient of 3% has been considered and, therefore, the force
transmitted to the abutment is 3% of the total permanent load of the deck.
For the force due to the temperature variation between rail and track an approx-
imate value of 2 MN has been adopted. An accurate value of this force due to the
via-structure interaction can be seen in the example in Chapter 6.

4.7.4.3 Forces Acting on the Abutment


In addition to these loads, the abutment itself is subjected to its own weight,
the weight of the earth gravitating on its underside, and the pressure of the soil
(Figure 4.22).

F2
10.90

F1

σ0 σ1

Figure 4.22 Scheme for the calculation of earth pressures.

@seismicisolation
4.7 Worked Example 139

Table 4.6 Vertical forces transmitted by the deck.

Vertical (MN)

Abutment self-weight 32.47


Friction slab self-weight 0.39 ⋅ L
Weight of soil on footing 21.35
Weight of soil on friction slab 3.07 ⋅ L

For the determination of the self-weight of the abutment, a specific weight of


25 kN/m3 and for the weight of the soil a value of 20 kN/m3 was considered.
In order to obtain these loads, the footing and the friction slab have been consid-
ered to have a width of 15.50 m, while for the body of the abutment a width of 14 m
has been considered.
Table 4.6 summarises the self-weight of the abutment, indicating separately the
self-weight of the friction slab, and the weight of the embankment gravitating on
the back of the footing and on the friction slab.
To determine the pressure of the earth on the abutment, the pressure at rest has
been used. The embankment head load is considered to be the track load and a
live load due to traffic, which together equate to a vertical pressure of 36 kN/m2
(Figure 4.22).
On the safety side, passive earth pressure has been considered: then formulae used
to calculate the pressures are:
k0 = 1 − sin φ (coefficient of passive earth pressure)

1 1
𝜎 1 = k0 γ h F1 = 𝜎 ⋅ b ⋅ h = γ ⋅ b ⋅ h2 ⋅ k0
2 1 2
𝜎 2 = k0 q F2 = k0 q ⋅ b h
where b = 14 m the width of the abutment, h = 10.90 the total height, g = 20 kN/m3
the weight of the soil embankment, and φ = 30 angle of internal friction of the soil.
The values obtained are:
F1 = 8.32 MN F2 = 2.73 MN F = F1 + F2 = 11.05 MN
Table 4.7 summarises the forces acting on the abutment.
The friction produced by the vertical loads acts as a stabilising force. A value of
0.45 has been adopted as the soil–slab friction coefficient, which is equivalent to an
angle of 24.2∘ .
If a sliding safety coefficient of 1.35 is assumed, the value of the minimum length
of the friction slab can be obtained from the following expression:
0.45 (63.49 + 3.46L)
1.35 =
25.92
And therefore the value of the length of the friction slab L will be 4.15 m. Adopting
a value of L = 4.50 m.
140 4 Design Basis

Table 4.7 Table Forces on abutments.

Vertical load (MN) Horizontal load (MN)

Abutment self-weight 32.47


Friction slab self-weight 0.39 ⋅ L
Weight of soil on footing 21.35
Weight of soil on friction slab 3.07 ⋅ L
Vertical load from deck 9.67
Earth pressure forces 11.05
Friction permanent loads 5.24
Braking and Traction 8.46
Temperature 𝛹 0 = 0.60 1.20
Total forces 63.49 + 3.46 ⋅ L 25.92

References
1 EN 1991-1-1 Eurocode 1: Actions on structures. Part 1.1: General actions.
2 EN 1991-2 Eurocode 1: Actions on structures. Part 2: Traffic loads on bridges.
3 EN 1337-1:2001 Structural bearings – Part 1: General design rules.
4 EN 1990 Eurocode 0: Basis of structural design.
5 EN 1991 Eurocode 1: Actions on structures – Part 2: Traffic loads on bridges, 2
Annex D Basis for fatigue assessment of railway bridges.
143

Dynamic Behaviour of High-Speed Railway Bridges


Alejandro Pérez-Caldentey

5.1 Introduction
Railroad loads are highly dynamic, and their effects must be determined accounting
for this fact. Isolated dynamic loads will produce deflections which can be higher
or lower than the maximum value of the load applied statically. Whether they are
higher or lower will depend on the variation of the load with time and on the ratio
between the load duration, td , and the structure’s first natural period of vibration, T.
The ratio between the dynamic and static deflection is known in literature as the
Dynamic Load Factor or DLF (also denoted as Φ). As a maximum, an isolated
load can produce a deflection equivalent to twice the maximum value of the load
(DLF = 2.0). For example, for a suddenly applied rectangular load, this value is
reached when the ratio td /T exceeds 0.5.
But, for the same type of loading, the deflection is only 60% of the static deflection
when td /T is equal to 0.1. The effect of other load variations with time will result in
smaller deflections. For an introduction to the dynamic response to moving loads in
SDOF systems please refer to Appendix A.
In railway bridges, the problem is complicated by the fact that the passage of a
train over a bridge is equivalent to applying a series of consecutive, periodic, isolated
loads. In such cases, resonance phenomena can occur. In a system without damping,
the response of an infinitely elastic structure can be unbounded (DLF = ∞) if the
period of the load coincides with a natural vibration period of the structure. For an
introduction to resonance effects please refer to Appendix A.

5.1.1 Resonance
For a railway bridge, the load is periodic due to the number of axles that go through
a bridge at a constant velocity. Referring to Figure 5.1, the period or frequency of the

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
144 5 Dynamic Behaviour of High-Speed Railway Bridges

d d d d d d
D D

OO OO OO OO OO OO OO OO OO OO
P P1 P2 P2 P2 P2 P1 P
D D D D D D D

Figure 5.1 Typical train definition.

Table 5.1 Resonance velocity for


extreme cases in km/h.

D (m)

T (s) 18 27

0.1 648 972

0.3 216 324

action can be determined from the velocity, by accounting for the spacing between
axle pairs (or bogies):
2π D D
Ti = =i i = 1,2, 3 … → vres,min = (5.1)
𝜔i v Tstr
The first natural period of a typical concrete high-speed viaduct can be in the
range of 0.1 to 0.3 seconds. The spacing, D, between pairs of axles of typical trains
varies between 18 and 27 m for trains with bogies. Using extreme values of the
above parameters, the velocity for which resonance occurs can be estimated. This
is done in Table 5.1. The smaller resonance velocity occurs for more flexible decks
and smaller distances between axle pairs. These numbers help to explain why
resonance is not normally considered for velocities below 200 km/h, as detailed
in the following sections. There is also no experimental evidence of resonance
problems in railway bridges for speeds lower than 200 km/h.

5.1.2 Envelope Dynamic Factor


The envelope dynamic factor is used when there is no risk of determining or critical
resonant phenomena and when the infrastructures being designed correspond to a
railway line along which very different type of convoys can circulate, so that it is
not practical to account for all the individual trains. For continuous structures, the
method may be applied when the design maximum velocity, v, is less than or equal to
200 km/h and the first natural frequency in bending of the structure, obtained con-
sidering only the permanent loads, n0 (Hz), falls within the limits given by Eq. (5.2):
n0,max [Hz] = 94.76L−0.748
𝛷
⎧ 80
⎪ for 4 m ≤ L𝛷 ≤ 20 m
n0,min [Hz] = ⎨ L𝛷 (5.2)
⎪23.58L−0.592 for 20 m ≤ L ≤ 100 m
⎩ 𝛷 𝛷
5.1 Introduction 145

L𝛷 is the determinant length in meters which depends on the element considered


and is given for a series of specific cases in EN 1991-2 Table 6.2 [1]. For the main
girders it is the span length for a simply supported beam and, for continuous spans,
a coefficient varying from 1.2 to 1.5 (depending on the number of continuous spans)
times the mean span length.
For continuous structures, even if the natural frequency does not fall within the
limits of Eq. (5.2), the formulation can be applied provided the first torsional fre-
quency is at least 20% larger than n0 and the ratio of nominal velocity to bending
frequency v/n0 complies with certain limits which are a function of the mass of the
structure, the span, and the damping index being considered. These limits are given
in Annex F of EN 1991-2.
For simple structures (i.e. ‘simply supported bridges with only longitudinal line
beam or simple plate behaviour with negligible skew effects on rigid supports’ [1])
and velocities larger than 200 km/h, the method can also be applied provided the first
torsional frequency is at least 20% larger than n0 and the ratio of nominal velocity to
bending frequency v/n0 complies with the limits given in Annex F of EN 1991-2.
Finally, simple structures with spans longer that 4000 m with train speeds greater
than 200 km/h are treated in the same way as continuous structures with speeds
smaller than 200 km/h.
When the conditions established in the previous paragraphs are fulfilled, a static
analysis is sufficient and the Dynamic Load Factor, which EN 1991-2, refers to sim-
ply as dynamic factor and denominates with the letter 𝛷, is applied to the design
train loads LM71, SW/0 or SW/2 to account for dynamic effects. This procedure
is assumed to provide an upper envelope covering all traffic on the bridge. In EN
1991-2, the value of the dynamic factor depends on the state of conservation of the
track and is given by the expressions of Eq. (5.3), where L𝛷 is in meters:
For a carefully maintained track
1.44
𝛷 = 𝛷2 = √ + 0.82 1.0 ≤ 𝛷2 ≤ 1.67
L𝛷 [m] − 0.2
For a track with standard maintenance
2.16
𝛷 = 𝛷3 = √ + 0.73 1.0 ≤ 𝛷3 ≤ 2.00 (5.3)
L𝛷 [m] − 0.2
From the upper limits that apply to the values of 𝛷, (i.e., ≤ 2.0) this factor does
not consider resonance and corresponds, as expected, to the DLF discussed in the
introduction to this section and in Appendix A.

5.1.3 Dynamic Factor for Real Trains Obtained by Means of Analytical


Formulations
The International Union of Railways publication UIC 776-1 [2] provides an expres-
sion for the DLF when real trains are considered. This formulation is also given in
EN 1991-2 Annex C. This formulation is useful when designing structures on which
a particular type of train will be circulating such as is the case with urban metro
systems. In this case the real loads of the vehicles can be used to assess structural
146 5 Dynamic Behaviour of High-Speed Railway Bridges

behaviour. The validity of the formulation is limited to structures having a natural


frequency greater than the lower limit given by Eq. (5.4) and maximum velocities
smaller than 200 km/h.
⎧ 80
⎪ for 4 m ≤ L𝛷 ≤ 20 m
n0 [Hz] ≥ ⎨ L𝛷 (5.4)
⎪23.58L−0.592 for 20 m ≤ L ≤ 100 m
⎩ 𝛷 𝛷

The dynamic factor for real trains is divided into the sum of the two factors, one if
the track has perfect geometry, 𝜑′ , and a second factor to account for imperfections
in the track, 𝜑′′ .
𝜑′ is given as a function of K, which depends on the first natural frequency of the
structure, n0 , the determinant length, L𝛷 , and the maximum nominal velocity, v, of
the real train:
v
K=
2L𝛷 n0
⎧ K
⎪ 4
if K < 0.76
𝜑′ = ⎨ 1 − K + K (5.5)
⎪1.325 if K ≥ 0.76

Figure 5.2 represents the value of 𝜑′ as a function of K. Its maximum value is equal
to 1.325.
The additional dynamic effect due to the irregularities of the track is given by
( )
min
v[m∕s] [ ( )
L [m] 2
( ) ( L [m] )2 ]
22 − 𝛷10 L𝛷 [m]n0 [Hz] − 𝛷20
𝜑′′ = 56e + 50 −1 e (5.6)
100 80

2.00

1.80

1.60

1.40

1.20

1.00
φ′

0.80

0.60

0.40

0.20

0.00
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
K

Figure 5.2 𝜑′ as a function of K.


5.1 Introduction 147

1.00 40

0.90
35
0.80
30
0.70
25
0.60

n0 (Hz)
φ″

0.50 20

0.40
15
0.30
10
0.20
5
0.10

0.00 0
0 10 20 30 40 50 60 70 80 90 100
LΦ (m)
φ″ for Upper limit for frequency φ″ for lower limit for frequency
n0 Upper limit n0 Lower limit

Figure 5.3 𝜑′′ as a function of the determinant length L𝜱 (for v > 80 km/h).

Figure 5.3 shows how 𝜑′′ varies with the determinant length for the upper and
lower limit natural frequencies for a maximum velocity greater than 80 km/h
(22 m/s). For smaller velocities, the values should be multiplied by a factor of
v[m/s]/22. This component of the dynamic factor is most important for bridges
with small spans and for secondary bridge components.
The total dynamic factor is a function of the type of rail track maintenance. The
static load due to railway traffic should be multiplied by 1 + 𝜑′ + 0.5𝜑′′ for a carefully
maintained track and by 1 + 𝜑′ + 𝜑′′ for standard maintenance.
The total dynamic factor for real trains is given in Figure 5.4 as a function of the
determinant length for the upper and lower natural frequency limits and for stan-
dard and careful track maintenance. The higher values of the dynamic factor apply
to the smaller determinant length. For standard maintenance and the lower limit of
the natural frequency the dynamic factor provides values which are slightly higher
than 2.00.

5.1.4 Dynamic Factor Obtained by Dynamic Analysis


When the limits to the analysis with a dynamic factor obtained from simple ana-
lytical equations as presented in Sections 5.1.2 or 5.1.3 are not fulfilled, then, it is
necessary to carry out a full dynamic analysis. This happens most notoriously for
continuous structures on high-speed railway lines, in which the 200 km/h speed
limit is exceeded, and therefore resonant phenomena are liable to occur, but it can
also be necessary for simply supported structures when the natural frequency of the
structure falls outside the limits established in Eq. (5.2).
The purpose of dynamic analysis is two-fold. On the one hand, it is aimed at iden-
tifying the velocity at which resonant frequencies occur and determine a dynamic
148 5 Dynamic Behaviour of High-Speed Railway Bridges

2.00

1.80

1.60
1 + φ′+ (1.00 or 0.5)φ″

1.40

1.20

1.00

0.80

0.60

0.40

0.20

0.00
0 10 20 30 40 50 60 70 80 90 100
LΦ (m)
n0,max Standard track maintenance n0,min Standard track maintenance
n0,max Careful track maintenance n0,min Careful track maintenance

Figure 5.4 Total dynamic factor as a function of the determinant length for the upper and
lower limits of the natural frequency and for careful and standard track maintenance.

load factor to be applied to the real trains to assess deflections and forces on the struc-
ture. The structure is to be designed for the worst case of the critical train with its
impact factor which can now go over the value of 200 due to resonance, or the design
train (LM71, SW/0 or SW/2) increased by the envelope dynamic factor, as outlined
in Section 5.1.2. On the other hand, dynamic analysis is carried out to determine
the maximum accelerations which are used to assess problems of track instability,
deconsolidation of ballast, and passenger comfort.
Dynamic analysis is to be carried out for a selection of real trains which should
account for ‘each permitted or envisaged train formation for every type of high-speed
train permitted or envisaged to use the structure at velocities over 200 km/h’ [1].
In order to cover this range for international lines where European high-speed
interoperability criteria are applicable, EN 1991-2 proposes two families of trains:
HSLM-A and HSLM-B. The second family is only relevant for simply supported
structures with span lengths less than or equal to 700 m. The HSLM-A family
comprises 10 trains which have different numbers of intermediate coaches (N),
different coach lengths (D), different axle bogie spacings (d), and different axle loads
(see Figure 5.5 and Table 5.2).
The mass of the system should consider all permanent loads (self-weight plus
removable loads: ballast, track, communications, handrails, etc.), obtained using
nominal values of density. An underestimation of the mass will lead to locate the
resonance point at larger velocities, since the resonance velocity is inversely propor-
tional to the period of the structure (see Eq. (5.1)) and the period becomes smaller
when the mass is smaller. On the other hand, an overestimation of the mass, which
leads to greater inertia, will lead to reduced maximum accelerations. Because of this,
5.1 Introduction 149

D N×D D

4×P 3×P 2×P 2×P 3×P 4×P

3 11 3 d d d d d d d 3 11 3
3.525 3.525

Figure 5.5 Geometry of HSLM-A trains.

Table 5.2 Definition of parameters of HSLM-A trains.

Number of Coach Bogie axle


Universal intermediate length spacing Point force
train coaches N D (m) d (m) P (kN)

A1 18 18 2.0 170
A2 17 19 3.5 200
A3 16 20 2.0 180
A4 15 21 3.0 190
A5 14 22 2.0 170
A6 13 23 2.0 180
A7 13 24 2.0 190
A8 12 25 2.5 190
A9 11 26 2.0 210
A10 11 27 2.0 210

two different assumptions regarding the permanent mass should be made when car-
rying out the dynamic analysis:
– a lower bound, designed to predict maximum deck accelerations, for which the
minimum ballast thickness and minimum ballast dry density (∼1700 kg/m3 )
should be used
– an upper bound, designed to determine the velocity at which resonant phenomena
occur using the maximum saturated density of ballast and accounting for future
track lifts.
With respect to stiffness, an overestimation of the stiffness will again lead to
an underestimation of the period of the structure and therefore (Eq. (5.1)) to an
overestimation of the resonance velocity. For this reason, the estimation of the
stiffness should tend to be a lower bound. Any potential cracking of concrete should,
therefore, be accounted for.
Regarding damping, the response of the structure in the range close to resonance
is sensitive to this value. EN-1991-2 Table 6.6 specifies the values of the damping
index to be used as a function of the material used for the deck of the structure (steel,
prestressed concrete, or reinforced concrete) and as a function of the span, as shown
in Table 5.3.
For bridges with two tracks, the dynamic effects should be estimated considering
only one track loaded.
150 5 Dynamic Behaviour of High-Speed Railway Bridges

Table 5.3 Damping index to be used for dynamic analysis (acc. to EN 1991-2 Table 6.6).

𝝃 – lower limit of percentage of critical damping (%)

Bridge Type Span L < 20 m Span L ≥ 20 m

Steel and composite 𝜉 = 0.5 + 0.125(20-L) 𝜉 = 0.5


Prestressed concrete 𝜉 = 1.0 + 0.07(20-L) 𝜉 = 1.0
Filler beam and reinforced concrete 𝜉 = 1.5 + 0.07(20-L) 𝜉 = 1.5

Dynamic analysis should be carried out for all HSLM trains plus real trains, if
required, at velocities going from 200 km/h to 1.2 times the maximum design speed
of the railway line. The velocity increment should not exceed 10 km/h.
The results of the dynamic analysis can be plotted in terms of maximum displace-
ment (see Figure 5.6) and maximum acceleration (see Figure 5.7) as a function of
the train velocity for each of the universal and real trains considered.
These figures are drafted for a simply supported prestressed concrete bridge with
a span, L, of 4500 m with a cross section having the following properties:
– Area, A = 7.96 m2
– Inertia, I y = 7.71 m4
– Modulus of elasticity of concrete (C45/55) Ec = 36.3 GPa.
Regarding the loads, the upper-bound weight due to ballast, track, communica-
tions, and handrails amounts to G2,sup = 137.6 kN/m (to be used for analysis of res-
onance effects), while the lower bound is G2,inf = 74.4 kN/m (to be used for the
determination of maximum accelerations).
It is interesting to estimate the first natural period of vibration of the structure
(T str ) and from its value infer the velocity for which resonance occurs by application
of Eq. (5.1). The period can be obtained by assimilation of the simply supported beam
to a SDOF system as outlined in [3]. By equating maximum kinetic energy (Ek ) and
maximum potential energy of the SDOF system, the maximum velocity of the system
is equal to the circular frequency times the maximum deflection:

1 1 KE
K y 2
= ME ẏ max → ẏ max =
2
y = 𝜔ymax (5.7)
2 E max 2 ME max
The equivalent mass is determined by the condition that the kinetic energy (K E )
of the SDOF system and that of the actual beam be the same (see Eq. (5.8)). 𝜙 is the
form function which describes the ratio between the deflection at a given coordinate
x and the maximum deflection.
2
⎛ ⎞
⎜ ⎟
1 1
L L
⎜ y(x) ⎟
2
M (𝜔ymax ) = m(x)(𝜔y(x)) dx → ME =
2
m(x)⎜ ⎟ dx
2 E 2 ∫0 ∫0 ⎜ ymax ⎟
⎜ ⏟⏟⏟ ⎟
⎝ 𝜙 ⎠
L
→ ME = m(x)(𝜙(x))2 dx (5.8)
∫0
5.1 Introduction 151

18

16

14

12
ymax (mm)

10

0
0 50 100 150 200 250 300 350 400 450
Train velocity (km/h)
Train A1 Train A2 Train A3 Train A4 Train A5
Train A6 Train A7 Train A8 Train A9 Train A10

Figure 5.6 Maximum displacement as a function of train velocity of HSLM-A.

0.30

0.25
Deck acceleration (g)

0.20

0.15

0.10

0.05

0.00
0 50 100 150 200 250 300 350 400 450
Train velocity (km/h)
Train A1 Train A2 Train A3 Train A4 Train A5
Train A6 Train A7 Train A8 Train A9 Train A10

Figure 5.7 Maximum acceleration as a function of train velocity for the 10 train types of
HSLM-A.
152 5 Dynamic Behaviour of High-Speed Railway Bridges

In this example, the equivalent mass is determined in Eq. (5.9) by using a sinu-
soidal form function for 𝜙(x):
M = mL = A ⋅ 2.5 ⋅ L + G2 ⋅ L = (7.96 ⋅ 2.5 + 13.76) ⋅ 45 = 1514.7 ton
L L( )
π 2
ME = m(𝜙(x))2 dx = m sin x dx
∫0 ∫0 L
( 2π )
L 1 + cos L x L 1514.7
=m dx = m = = 757.35 ton (5.9)
∫0 2 2 2
The equivalent stiffness of the SDOF is obtained by dividing the equivalent
load, PE , by the maximum deflection ymax . The equivalent load is determined by
establishing that the external work done by the forces in the SDOF system is equal
to the external work done by the forces on the real system:
L L
y(x)
PE ymax = p(x)y(x)dx → PE = p(x) dx
∫0 ∫0 ymax
⏟⏟⏟
𝜙
L
PE
→ PE = p(x)𝜙(x)dx → KE = (5.10)
∫0 ymax
For the given example:
q(x) = q
L L ( ) [ ( )]L
π L π 2
PE = q(x)𝜙(x)dx = q sin x dx = q cos x = qL
∫0 ∫0 L π L 0 π
= 0.64 qL
PE 0.64qL EI 36.2 ⋅ 106 ⋅ 7.71
KE = = 4
= 49.152 3 = 49.152 ⋅
ymax 5 qL L 453
384 EI
= 150 545.1 kN∕m (5.11)
The first natural period would then be:
√ √
ME 735.35
T = 2π = 2π = 0.44 s (5.12)
KE 150 545.1
Given that the axle pair spacing of the Universal trains (which is equal to the coach
length, D) is between 18 and 27 m, it can be expected, by virtue of Eq. (5.1), that
resonant peaks would occur at velocities ranging from 18/0.44 ⋅ 3.6 = 147 km/h to
27/0.44 ⋅ 3.6 = 221 km/h. This is what can be observed in Figures 5.6 and 5.7. This
implies that it is not necessary to cover the full range of velocities when trying to
determine the maximum Dynamic Load Factor. It would be sufficient to verify the
resonant frequencies and their multiples, as well as the maximum velocity.
The effect of track irregularities can be accounted for by adding to the dynamic
coefficient determined by dynamic analysis 0.5 𝜑′′ for careful track maintenance or
𝜑′′ for standard track maintenance. High-speed railways always meet careful main-
tenance criteria.
5.2 Methods for Dynamic Calculations and Structural Response 153

5.2 Methods for Dynamic Calculations and Structural


Response

5.2.1 Modal Superposition


The modal superposition method can be formulated for finite elements as a function
of the degrees of freedom of the structure. In this case, the eigenforms result from
the analysis. Another approach, valid for simple structures is to assume a certain
shape for the eigenmodes. The two approaches are developed below.

5.2.1.1 Matrix Formulation for Finite Element Analysis


The dynamic equilibrium of a system with multiple degrees of freedom (MDOF) can
be written in matrix form as expressed in Eq. (5.13).
Mÿ + Cẏ + Ky = F(t) (5.13)
By multiplying both sides of the equation by the inverse of the mass matrix M, this
equation becomes:
ÿ + M−𝟏 Cẏ + M−𝟏 Ky = M−𝟏 F(t) (5.14)
In most practical cases, M can be assumed to be a diagonal matrix which is formed
by the mass lumped at the nodes of the model. The values of the diagonal (i = j), M i ,
would be the translational or rotational mass lumped at the node corresponding to
degree of freedom i, with all other terms (i ≠ j) being equal to 0.
Modal superposition is a simplified technique to solve the dynamic analysis of
systems with n degrees of freedom (MDOF) by reducing them to solving a number
m (m ≤ n) of SDOF systems. This simplification is achieved by eigen-decomposition
of matrix M−1 K. 𝚽 will denote the matrix whose columns are formed by the
eigenmodes.
Eigenvectors are not unique and to avoid this indeterminacy, they can be nor-
malised, for instance, by imposing that their Euclidean norm is equal to 1.
The eigenvectors of a matrix form a base and y can be expressed in terms of this
base as shown in Eq. (5.15):
y = 𝚽 q (5.15)
⏟⏟⏟ ⏟⏟⏟⏟⏟⏟
n⋅1 n⋅m m⋅1
By introducing Eq. (5.15) into (5.13), the following expression is obtained:
M𝚽q̈ + C𝚽q̇ + K𝚽q = F(t) (5.16)
By multiplying Eq. (5.16) by the transpose of the eigenvalue matrix 𝚽T , Eq. (5.17)
is obtained:
𝚽T M𝚽 q̈ + 𝚽T C𝚽 q̇ + 𝚽T K𝚽 q = 𝚽T F(t) (5.17)
⏟⏟⏟ ⏟⏟⏟ ⏟⏟⏟
𝚲M =diag(Mi∗ ) 𝚲C =diag(Ci∗ ) 𝚲K =diag(Ki∗ )

It can be demonstrated that the eigenvectors of M−1 K are orthogonal with respect
to both M and K (for a demonstration see the Appendix of reference [3]). This makes
matrices 𝚽T M𝚽 and 𝚽T K𝚽 diagonal.
154 5 Dynamic Behaviour of High-Speed Railway Bridges

To decouple the system of equations it is also necessary that Matrix C be orthog-


onal with respect to 𝚽. This is possible by using a type of damping called Caughey
damping, for which the damping matrix is expressed as a sum of powers of M−1 K.

n
C=M ai (M−𝟏 K)i (5.18)
i=0

This is equivalent to having a damping index for mode i equal to:

1∑
n−1
𝜉i = a (𝜔 )2j−1 (5.19)
2 j=0 j i

The values ai can be determined for the system of Eq. (5.19), by imposing values
of the damping index for each vibration mode.
With this assumption Eq. (5.17) becomes a series of m independent SDOF differ-
ential equations, which are easy to solve. m is the number of eigenvalues considered
in the analysis. If the elements of diagonal matrix 𝚲K are named Mi∗ , the elements
of diagonal matrix 𝚲K are named Ki∗ and the elements of diagonal matrix 𝚲C , are
named Ci∗ , then the expression of Eq. (5.20) holds:

Ci∗ Ki∗ ∑
n
𝜙−1 ij
q̈ i + ̇
∗ qi + ∗ qi = Fj (t)jy
Mi Mi j=1
Mi∗
Ci∗ = 𝛟−𝟏
i C𝛟i
Mi∗ = 𝛟−𝟏
i M𝛟i
Ki∗ = 𝛟−𝟏
i K𝛟i (5.20)

In Eq. (5.20), jy is a directional coefficient which is equal to 1.00 if force F j (t) goes
in the direction of degree of freedom i and is equal to 0.00 if it does not.
The forces applied on the nodes F j (t) can be simulated by triangular pulses
(Figure 5.8). The time between pulses can be approximated as the spacing between
bogies, D, divided by the velocity of the train, v. The duration of the triangular pulse
would be equal to the sum of the distances to the adjacent nodes divided by the
velocity.
Once the independent SDOF systems have been solved, their effects need to be
superimposed. This can be done, of course by direct summation of the displace-
ments, or acceleration time histories. However, this procedure is time consuming.
For this reason, it is common practice to determine the maximum response of each
vibration mode at a given location and superimpose the maximum effect by using
a combination rule that considers that the maximum response from the different
vibration modes is not likely to occur simultaneously. A classical combination rule
is the Square Root of the Sum of the Squares (SRSS), given in Eq. (5.21):

√m
√∑
E = √ Ej2 (5.21)
j=1
5.2 Methods for Dynamic Calculations and Structural Response 155

450
Fj (t) Load applied on a given node (kN)

400 D/v

350

300

250

200

150

100

50 (li–1 + li)/v

0
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80 2.00
t (s)
Load vs time on a node

Figure 5.8 Triangular pulses simulating the train loads.

where:

E is the total estimated maximum response (displacement, acceleration, velocity) of


the structure
Ej is the maximum response (displacement, acceleration, velocity) of vibration
mode j.

This criterion can, however, be unsafe when the periods of the vibration modes
differ by less than 10%. In such cases the Complete Quadratic Combination (CQC)
should be applied. This combination criterion is defined in Eq. (5.22):

√m m √
√∑∑
E=√ Ek rkj Ej = ET rE
k=1 j=1
√ ( ) 3∕2
8 𝜉k 𝜉j 𝜉k + 𝜌kj 𝜉j 𝜌kj
rkj = ( )2 ( ) ( )
1 − 𝜌2kj + 4𝜉k 𝜉j 𝜌kj 1 + 𝜌2kj + 4 𝜉k2 + 𝜉j2 𝜌2kj
Tk
𝜌kj = ≤1 (5.22)
Tj
This expression is very general and accounts for the possibility of having different
damping indexes for the different vibration modes.

5.2.1.2 Formulation Based on Assumed Eigenforms


The formulation of Section 5.2.1.1 is the general formulation used in finite element
analysis. It is also possible to formulate the problem by assuming a certain shape for
the eigenforms, respecting the boundary conditions. For a simply supported beam,
156 5 Dynamic Behaviour of High-Speed Railway Bridges

1.50

1.00

0.50

0.00
Φ(x)

0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00

–0.50

–1.00

–1.50
x/L
n=1 n=2 n=3 n=4

Figure 5.9 Sinusoidal eigenforms for a simply supported beam.

for example, a sine function can be adequately used to model the eigenforms (see
Eq. (5.23) and Figure 5.9):
π
𝜙(x) = 𝜙0 sin n x (5.23)
L
To maintain the criterion that the norm of the eigenforms is equal to 1.00, 𝜙0
should be fixed as follows:
L
∫0 𝜙2 (x)dx 𝜙20 ( L ) 𝜙2 L ( ( ))
π π
= 1.00 = sin2 n x dx = 0 1 − cos 2n x dx
L L ∫0 L 2L ∫0 L
𝜙20 [ L
(
π
)]L 𝜙2 √
= x− sin 2n x = 0 = 1.00 → 𝜙0 = 2 (5.24)
2L 2nπ L 0 2
In general, to satisfy varying boundary conditions (fixed displacements or rota-
tions), the eigenforms can take on the shape shown in Eq. (5.25) [3]:
nπ nπ nπ nπ
𝜙(x) = a cos x + b sin x + c cosh x + d sinh x
L L L L
𝜕 4 𝜙(x) n4 π4
= 𝜙(x) (5.25)
𝜕x4 L4
⏟⏟⏟
(an )4

Note the property shown in the second line of the equation (i.e. that the fourth
derivative of the eigenform with respect to x is equal to the eigenform times the
fourth power of the form coefficient an ).
An expression for the natural frequencies of the system can be determined from
dynamic equilibrium conditions. It is well known from the static sectional equilib-
rium equation of moments that the moment is proportional to the curvature and the
5.2 Methods for Dynamic Calculations and Structural Response 157

mдxÿ

M + дM V V M V V + дV

дx дx

Figure 5.10 Equilibrium of a beam slice.

proportionality constant is the flexural stiffness of the section:


𝜕2 y
M = −EI (5.26)
𝜕x2
From the equilibrium conditions of a slice of a beam, it can be established that the
shear force is the derivative of the bending moment and that the load per meter, p,
applied on the slice minus the inertial forces is equal to the derivative of the shear
force (see Figure 5.10).
𝜕M 𝜕3 y
V= = −EI 3
𝜕x 𝜕x
𝜕V 𝜕4 y
p − m̈y = − = EI 4 (5.27)
𝜕x 𝜕x
Developing the second expression of Eq. (5.27), and assuming that the slice is
vibrating freely, so that p = 0:
𝜕4 y
m̈y + EI = p(x, t) = 0⎫
𝜕x4 ⎪

y = 𝜙(x)q(t) ⎪
⎪ 𝜕4 𝜙
𝜕 4𝜙
⎬ → −m𝜙𝜔 q + EI 4 q = 0
2
m𝜙q̈ + EI 4 q = 0 𝜕x
𝜕x ⎪

q(t) = a cos 𝜔t + b sin 𝜔t ⎪

̈ = −𝜔2 q(t)
q(t) ⎭

EI 1 𝜕 4𝜙
EI 1 𝜕 4 𝜙
→ 𝜔2 = → 𝜔 = (5.28)
m 𝜙 𝜕x4 m 𝜙 𝜕x4
With this approximation – i.e. assuming a shape for the eigenmodes – Eq. (5.20)
can be reformulated as follows:
L
C∗ K∗ 1
q̈ n + n∗ q̇ n + n∗ qn = ∗ F (x, t)𝜙n (x)dx
Mn Mn Mn ∫0 j
L
Cn∗ = c (𝜙n (x))2 dx = cL
∫0
158 5 Dynamic Behaviour of High-Speed Railway Bridges

L
Mn∗ = m (𝜙n (x))2 dx = mL
∫0

𝜕 4 𝜙 EI L 2 𝜕 4 𝜙 EI
Kn∗ = (𝜙 (x)) dx = L
𝜕x4 𝜙n ∫0 n
𝜕x4 𝜙n
L
c 𝜕 4 𝜙 EI 1
→ q̈ n + q̇ n + 4 qn = F (x, t)𝜙n (x)dx (5.29)
m 𝜕x 𝜙n m mL ∫0 j
⏟⏞⏞⏟⏞⏞⏟
𝜔2

For the case of a simply supported beam, subjected to a given generic load
F max f (𝜏), the application of Eq. (5.29) would result in:

𝜙n = 𝜙0 sin x
L
𝜕4 𝜙 n4 π4 nπ n4 π4
= 𝜙0 4 sin x = 4 𝜙n (x)
𝜕x 4 L L L
y(t) = 𝜙n (x)qn
L
c n4 π4 EI F
q̈ n + q̇ n + 4 qn = max f (𝜏, x)𝜙n (x)dx (5.30)
m L m mL ∫0
⏟⏟⏟
𝜔2

To obtain the deflection at a given abscissa, y(x), the value of q(x) must be multi-
plied by the value of the eigenform at that coordinate so that:
nπx
y(x) = q(x)𝜙0 sin (5.31)
L
Applications of this equation are given in Sections 5.2.2 and 5.2.3.

5.2.2 Response to the Isolated Load


It is straightforward to apply the formulation of Eq. (5.30) to the case of a simply
supported beam subjected to a moving load. The dynamic equilibrium for the time
comprised between the moment the load enters the bridge until it exits the bridge
is given by Eq. (5.32). Once the load exits the bridge, the solution will be that of a
damped harmonic system subjected to the initial position and velocity conditions
present at the time the load exits the bridge.
L
c n4 π4 EI P
q̈ n (x) + q̇ n (x) + 4 q (x) = f (𝜏, x)𝜙n (x)dx
m L m n mL ∫0
⏟⏟⏟
𝜔2
L
f (𝜏, x) = 𝛿(x − vt) → 𝛿(x − vt)𝜙n (x)dx = 𝜙n (vt)
∫0
⎛ ⎞
⎜ ⎟
c P ⎜ nπ ⎟
q̈ n (x) + q̇ n (x) + 𝜔 qn (x) =
2
𝜙 sin ⎜
L ⎟⎟
vt (5.32)
m mL 0 ⎜⏟⏟ ⏟
⎜ ⎟
⎝ 𝜔f ⎠
5.2 Methods for Dynamic Calculations and Structural Response 159

𝛿(x) in Eq. (5.32) is Dirac’s delta function which is equal to infinity when x = 0 and
equal to 0 otherwise. It has the property shown in Eq. (5.33):

𝛿(x − a)f (x)dx = f (a) (5.33)



The problem of Eq. (5.32) is in fact the same problem of forced vibrations solved in
Appendix A. However, in this case, damping is accounted for. It can be verified that
the sum of a sine function plus a cosine function with natural frequency equal to
𝜔f is a particular solution of Eq. (5.32). The full solution to the equation will be the
sum of the particular solution and the general solution to the homogenous equation
(damping is ignored here for the part of the solution corresponding to forced vibra-
tion, since it will be neglectable, for typical bridge damping ratios, because of its
short duration):
q = A cos 𝜔f t + B sin 𝜔f t + e−𝜉𝜔t (C cos 𝜔d t + D sin 𝜔d t)
q̇ = −A𝜔f sin 𝜔f t + B𝜔f cos 𝜔f t − 𝜉𝜔e−𝜉𝜔t (C cos 𝜔d t + D sin 𝜔d t)
+ 𝜔d e−𝜉𝜔t (−C sin 𝜔d t + D cos 𝜔d t)
q̈ = −A𝜔2f cos 𝜔f t − B𝜔2f sin 𝜔f t + 𝜉 2 𝜔2 e−𝜉𝜔t (C cos 𝜔d t + D sin 𝜔d t)
− 𝜉𝜔d 𝜔e−𝜉𝜔t (−C sin 𝜔d t + D cos 𝜔d t) − 𝜉𝜔d 𝜔e−𝜉𝜔t (−C sin 𝜔d t + D cos 𝜔d t)
− 𝜔2d e−𝜉𝜔t (C cos 𝜔d t + D sin 𝜔d t)
P
q̈ n (x) + 2𝜉𝜔q̇ n (x) + 𝜔2 qn (x) = 𝜙0 sin 𝜔f t (5.34)
m
The values of A and B can be identified by imposing that the terms in cos𝜔f t and
sin𝜔f t must cancel out, as shown in Eq. (5.35)

⎧ P −2𝜉𝜔𝜔f
⎧ 2𝜉m𝜔
⎪A = 𝜙0 m ( )2
⎪ ( ⏞⏞⏞ ⎪ 𝜔2 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f
)
⎪ A 𝜔2 − 𝜔2 + B c 𝜔 = 0 ⎪ ( )
⎨ f m f →⎨ (5.35)
⎪ ( ) ⎪ 𝜔2 − 𝜔2f
P
⎪−A c 𝜔 + B 𝜔2 − 𝜔2 = 𝜙 P ⎪ B = 𝜙0 ( )2
⎩ m f f 0
m ⎪ m
⎩ 𝜔2 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f

The values of coefficients C and D can then be determined from the initial condi-
tions of a system that is initially at rest:
−2𝜉𝜔𝜔f 𝜙0 mP
q(0) = 0 = A + C = ( )2 +C
𝜔 − 𝜔f + 4𝜉 𝜔 𝜔f
2 2 2 2 2

2𝜉𝜔𝜔f 𝜙0 mP
→C= ( )2
𝜔2 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f

P 2𝜉𝜔𝜔f
→ C = 𝜙0 ( )2
m
𝜔2 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f
̇
q(0) = 0 = B𝜔f − 𝜉𝜔C + 𝜔d D
160 5 Dynamic Behaviour of High-Speed Railway Bridges

( )
𝜔2 − 𝜔2f 𝜙0 mP 𝜔f 2𝜉𝜔𝜔f 𝜙0 mP
𝜔
→ D = −( )2 𝜉
𝜔d ( 2 )2
+
𝜔
𝜔2 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f d 𝜔 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f
𝜔f
( ( ))
2𝜉 2 𝜔2 − 𝜔2 − 𝜔2
P 𝜔d f
→ D = 𝜙0 (5.36)
m ( 2 )2
𝜔 − 𝜔2f + 4𝜉 2 𝜔2 𝜔2f

The position and velocity of the structure can be determined when the force exits
the structure as follows:
𝜙0 P
q0 = ( )2
m 𝜔2 − 𝜔2f + 4𝜉𝜔2 𝜔2f
( )
⎛−2𝜉𝜔𝜔f cos 𝜔f L + 𝜔2 − 𝜔2f sin 𝜔f L ⎞
⎜ v v ⎟
⋅⎜ ( ( ) )⎟
𝜔
⎜+e−𝜉𝜔 v 2𝜉𝜔𝜔f cos 𝜔d L + f 𝜔2 − 𝜔2 sin 𝜔f L ⎟
L

⎝ v 𝜔d f v ⎠
𝜙0 P
q̇ 0 = ( )2
m 𝜔2 − 𝜔2f + 4𝜉𝜔2 𝜔2f
( )
⎛2𝜉𝜔𝜔2 sin 𝜔f L + 𝜔f 𝜔2 − 𝜔2f cos 𝜔f L ⎞
⎜ f v v ⎟
⎜ L ⎟
⎜ ⎛2𝜉𝜔𝜔f cos 𝜔d ⎞⎟
⎜−𝜉𝜔e−𝜉𝜔 Lv ⎜ v ⎟⎟
⎜ ⎜ 𝜔f ( ( ))
L ⎟⎟
⋅⎜ ⎜+ 2𝜉 2 𝜔2 − 𝜔2 − 𝜔2f sin 𝜔d ⎟ ⎟ (5.37)
⎜ ⎝ 𝜔d v ⎠⎟
⎜ ⎛−2𝜉𝜔𝜔f sin 𝜔d L ⎞⎟
⎜ −𝜉𝜔 Lv ⎜ v ⎟⎟
⎜+𝜔d e ⎜ 𝜔f ( ( ))
L⎟

⎜ ⎜+ 2𝜉 2 𝜔2 − 𝜔2 − 𝜔2f cos 𝜔d ⎟⎟
⎝ ⎝ 𝜔d v ⎠⎠
The above equations can be applied for the number of eigenmodes considered. To
obtain the actual displacements, the generalised coordinates q, should be multiplied
by the value of the eigenmode at the given location (see Eq. (5.31)). Therefore, the
contribution to the deflection at centre span for pair values of n will be nil.
To see an application of this method, the same example already analysed in
Section 5.1.4 is considered. The example consists in a simply supported bridge with
a span, L, of 4500 m. The area of the section is 7.96 m2 and the inertia 7.71 m4 . The
elastic modulus of concrete is taken as 36.2 GPa. Additionally, an upper-bound
superimposed dead load equivalent to G2,sup = 137.6 kN/m is present. So, the total
upper-bound mass of the structure is M = (7.96 ⋅ 2.5 + 13.76) ⋅ 45 = 1514.7 ton. The
first natural period of this structure is T1 = 0.448 s. It is assumed that the train is a
high-speed train, and that the maximum velocity is 350 km/h (i.e., v ∼ 100 m/s) and
that the damping index is 1% as recommended by EN 1991-2 (see Table 5.3).
Figure 5.11 shows the response of the system in terms of maximum deflection
at centre span for the first four modes of vibration. It is very clear that in such a
5.2 Methods for Dynamic Calculations and Structural Response 161

340 kN loads travelling over a 45 m simply supported span at 350 km/h


Deflection at centre span for modes 1 through 4
5.00

4.00 0.03

3.00
0.02

y2, y3, y4 (mm)


2.00
0.01
1.00
y1 (mm)

0.00 0
0.00 0.50 1.00 1.50 2.00 2.50 3.00
–1.00
–0.01
–2.00
–0.02
–3.00

–4.00 –0.03
Time (s)

1st mode: n = 1 2nd mode: n = 2


3rd mode: n = 3 4th mode: n = 4

Figure 5.11 Deflections due to the first four vibration modes – plotted separately.

system, only the first mode of vibration is of significance (note that the deflection of
modes 2, 3, and 4 is plotted on the secondary axis on a different scale, roughly 100
times larger). Modes with pair values of n do not contribute at all to the deflection
at centre span since the deflection of the corresponding eigenmodes is nil at that
point. Figure 5.12 shows the deflection due to the superposition of the first four
modes compared to that of the first model only. The maximum deflection of the
superposed modes is 0.64% lower than that of the first mode only and the difference
can barely be seen.
Table 5.4 shows the natural frequencies of the first four modes, the maximum
centre span deflection due to each mode, the deflection due to the sum of the modes
and the deflection obtained by the SRSS criterion, which would be applicable since
the difference between the periods of the first four modes is much larger than 10%.
The SRSS criterion is safe sided for this case since the contribution of the third mode
is negative at centre span.

Table 5.4 Maximum deflection in the structure [mm].

Mode 1 2 3 4

𝜔 (rad/s) 14.03 56.14 126.31 224.55


T (s) 0.448 0.112 0.050 0.028 Sum (mm) SRSS (mm)
Deflection (mm) 3.815 0.000 −0.029 0.000 3.790 3.815
162 5 Dynamic Behaviour of High-Speed Railway Bridges

340 kN loads travelling over a 45 m simply supported span at 350 km/h


5.00

4.00

3.00

2.00
y (mm)

1.00

0.00
0.00 0.50 1.00 1.50 2.00 2.50 3.00
–1.00

–2.00

–3.00

–4.00
Time (s)

1st mode: n = 1 Sum of first 4 modes

Figure 5.12 Deflections due to the additive effect of the first four vibration modes.

5.2.3 Response to the Train Loads


The response to the train load can be computed by superposing the effect of a
number N of axes with a time delay equal to the distance between bogies, D, divided
by the velocity of the train. In the case of the train load, the maximum response is
not obtained for the fastest train passage but for a time between loads equal to the
natural period of the structure. The following figures have been obtained by apply-
ing this methodology to the example considered in Section 5.1.4. In this case the
natural period is 0.448 s. 14 bogies carrying a total load of 340 kN each, and spaced
22 m apart, are traveling at a velocity equal to 22/0.448 = 49.1 m/s (176.8 km/h).
The damping index 𝜉 is assumed to be 1%. As a simplification, which is justified by
the results shown in Section 5.2.2, only the first vibration mode will be considered
here. Figure 5.13 shows the individual effect on the bridge deflections of the
passage of the first 4 bogies. The vibrations once the train leaves the structure are
in synchronicity.
Figure 5.14 shows the superposition of the passage 7 and 14 bogies, compared to
the effect of one bogie, in terms of maximum deflection. A clear resonant effect can
be seen. The deflection increases with time, until the last bogie leaves the deck, after
which moment the maximum deflection goes down as only the vibrations left over
from the passage of the 14 bogies are active. These vibrations gradually decrease due
to damping.
5.2 Methods for Dynamic Calculations and Structural Response 163

340 kN loads travelling over a 45 m simply supported span at 177 km/h


3.50

3.00

2.50

2.00

1.50
y (mm)

1.00

0.50

0.00
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00 4.50 5.00
–0.50

–1.00

–1.50
Time (s)
Bogie #1 Bogie #2 Bogie #3 Bogie #4

Figure 5.13 Effect of individual loads.

Superposition of 7 and 14 bogies compared with the effect of one


15.00

10.00

5.00
y (mm)

0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.00

–5.00

–10.00

–15.00
Time (s)
1 Bogie 7 Bogies 14 Bogies

Figure 5.14 Superposition of several bogies.


164 5 Dynamic Behaviour of High-Speed Railway Bridges

Dynamic Load Factor as a function of the train velocity


4.50

4.00

3.50

3.00

2.50
Φ

2.00

1.50

1.00

0.50

0.00
0 50 100 150 200 250 300 350
Train velocity, v (km/h)

Effect of 14 bogies Effect of 2 bogies

Figure 5.15 Dynamic Load Factor (𝛷) as a function of the train velocity.

Figure 5.15 shows the Dynamic Load Factor as a function of the train velocity.
A very clear resonant peak can be observed at 177 km/h (corresponding to the first
natural period of vibration).1 Also shown in the figure is the effect of two bogies
(which is the maximum number that will be on the bridge at a given time). With
only two bogies, there is no resonant phenomenon and, in this case the response
increases with the velocity of the train. This is logical as more energy is transferred
to the structure and the ratio between the load duration and the natural period of
the structure is not too low (td /T ∼ 0.5 for a velocity of 350 km/h).

5.2.4 Effect of Damping


In real structures damping values are low (going from 0.5–1% for steel decks to 2% for
concrete decks). These values are not very significant for individual loads. However,
when train load is considered, the time lapse between the passage of the successive
bogies provides some room for action by damping. Even with these low damping
indexes the effect is quite significant. Figure 5.16 shows how the Dynamic Load

1 It is interesting to note that for such a speed, a dynamic analysis would not be required by
EN 1991-2:2003, despite the fact that resonance occurs. In this case, the Dynamic Load Factor
estimated according to Eq. (5.3) would be only 1.03. However, this factor is applied to the LM71 or
SW/0, SW/2 trains which are much heavier than the real trains. For the LM71 train, the maximum
static deflection would be 18.6 mm, much larger than the 13.44 mm of the real train accounting for
the dynamic behaviour. This comparison demonstrates that such cases are covered by the envelope
loading.
5.2 Methods for Dynamic Calculations and Structural Response 165

Effect of damping on the Dynamic Load Factor


6.00

5.00

4.00

3.00
Φ

2.00

1.00

0.00
0 50 100 150 200 250 300 350
Train velocity, v (km/h)
0% damping 1% damping 2% damping

Figure 5.16 Effect of damping on Dynamic Load Factor.

Factor varies when there is no damping compared to 1% and 2% damping indexes.


For the peak response, 𝛷 increases from 3.2 for a 2% damping index to close to 5.6 if
damping is not accounted for.

5.2.5 Dynamic Interaction Between Vehicle and Structure


The interaction between vehicle and structure normally has a favourable effect on
the dynamic behaviour of railway bridges. According to [4], its effects are most
noticeable in simply supported bridges with small spans and low damping. In these
cases, the reduction in accelerations and displacements at resonance can reach 30%.
There are several levels for the modelling of the interaction between vehicle and
structure. The most sophisticated models consider not only the interaction between
vehicle and structure but also the interaction with the track, in which rails, sleep-
ers, and ballast are all considered (see Figure 5.17). Such models, however, are too
complex and are more useful for research than for practical applications.
A second-level approximation, which has been applied in practice, considers the
modelling of each vehicle with two bogies, and represents the connection between
car and bogie and the connection between bogie and each axle (see Figure 5.18).
In this case the body of the vehicle is assigned a mass (M) and a rotational inertia (J).
The vehicle is connected, though secondary suspension (modelled by a spring -K s -
plus a damper -Cs - placed in parallel), to the bogie frame which is also assigned a
mass (M B ) and a rotational inertia (J B ). Each axle is connected to the bogie frame by
166 5 Dynamic Behaviour of High-Speed Railway Bridges

Figure 5.17 Dynamic interaction model considering the interaction between vehicle,
track, and structure [5].

dtd dBd

Secondary
suspension M, J

MB, JB MB, JB

Primary
suspension

MW
LB deB
L

Figure 5.18 Full vehicle–structure interaction model (Source: taken from [4]).

Ms Figure 5.19 Simplified model to


account for vehicle–
structure interaction.

Kp Cp

Mu

secondary suspension modelled as a coupled spring/damper pair (K p /Cp ) for each


axle. Finally, the mass of the wheels (M w ) is applied directly on the structure.
There are also simplified models in which the axles are considered as indepen-
dent elements and only primary suspension is considered. Each one can be mod-
elled as shown in Figure 5.19. In this case M s corresponds to the suspended mass
5.3 Interoperability 167

(equivalent to the mass of the bogie supported by the axis plus the weight of vehicle
box supported by the axle), while M u corresponds to the mass of the wheels and the
axle itself. These models neglect the coupling provided by the vehicle box and the
bogies.

5.3 Interoperability

5.3.1 Introduction
The European Union has issued two directives and one amendment on the inter-
operability of European railways systems:
– Directive 96/48/EC on the interoperability of the trans-European high-speed
railway system
– Directive 2001/16/EC on the interoperability of the trans-European conventional
railway system
– Directive 2004/50/EC which amends the above.
The objective of the directives is to facilitate the circulation of railway traffic in the
different railway networks existing in Europe. To this end, the directives target the
design, construction, serviceability, and exploitation phases setting a set of common
standards.
Directive 96/48/EC defines interoperability as ‘the ability of the trans-European
high-speed rail system to allow the safe and uninterrupted movement of high-speed
trains which accomplish the specified levels of performance’. This objective is to be
achieved by meeting regulatory, technical, and operational conditions.
An important technical condition is related to the traffic loading. For this pur-
pose, EN 1991-2 defines a series of universal dynamic trains for which all bridges
of the European Networks must be designed. These are presented in Sections 5.3.2
and 5.3.3.

5.3.2 Universal Dynamic Train A


In order to cover the loading range for international lines where European
high-speed interoperability criteria are applicable, EN 1991-2 proposes two families
of trains: HSLM-A and HSLM-B. The HSLM-A family comprises 10 trains which
have different number of intermediate coaches (N), different coach lengths (D), dif-
ferent axle bogie spacings (d), and different axle loads (see Figure 5.5 and Table 5.2).

5.3.3 Universal Dynamic Train B


Family HSLM-B is only relevant for simply supported structures with spans length
less than or equal to 7.00 m. The HSLM-B train is formed by a number N, which
depends on the span, of 170 kN loads equally spaced at a distance d, which also
168 5 Dynamic Behaviour of High-Speed Railway Bridges

= = = = = = = = = d = = = = = = =

Figure 5.20 Geometry of HSLM-B trains.

4.65
3.35

3.8

4.2

4.8
3.5
6.00 20

5.50
Spacing between load, d (m)

4.35
3.65

5.65
2.8
3.2

4.5

5.8
5.00 15

4.50

6.15
2.65

5.15

5.5
4.00 10

N
4.35
1.2

4.5

5.8
3.65

3.50
4.2
1.6

3.5

3.00 5
4.65

5.65
2.50

2.00 0
0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2 4.8 5.4 6.0 6.6
Span, L (m)
d (m) N

Figure 5.21 Values of d and N as a function of the span.

depends on the span (see Figure 5.20). Figure 5.21 provides the values of these vari-
ables as a function of the span.

5.4 Application Examples


5.4.1 Case Without Dynamic Analysis
A three-span continuous structure with span distribution of 36-45-36 carries a rail-
way track with maximum velocity limited to 200 km/h. It is assumed that the track is
carefully maintained. The first natural period of vibration is T = 0.307 s. According
to Figure 6.9 of EN 1991-2:2003, if the bridge is continuous, then a dynamic analysis
is not required, if the first natural frequency of the structure n0 is within the limits
shown in Eq. (5.2). In this case the static analysis using the envelope dynamic fac-
tor is sufficient. The determinant length, L𝛷 for the global analysis of the structure
is obtained from Table 6.2 of EN 1991-2:2003. The determinant length is equal to a
factor k, which is equal to 1.3 for 3 spans times the mean span of the deck. In this
case: L𝛷 = 1.3 ⋅ 39 = 50.7 m.
5.4 Application Examples 169

The frequency limits of Eq. (5.2) for this case are:


n0,max = 94.76L−0.748
𝛷 = 94.76 ⋅ 50.7−0.748 = 5.03 Hz → T = 0.20 s
n0,min = 23.58L−0.592
𝛷 = 23.58 ⋅ 50.7−0.592 = 2.30 Hz → T = 0.43 s (5.38)
The structure’s first frequency complies with the limits required for the application
of the envelope dynamic load factor.
For a carefully maintained track:
1.44 1.44
𝛷= √ + 0.82 = √ + 0.82 = 1.028 1.0 ≤ 𝛷 ≤ 1.67 (5.39)
L𝛷 − 0.2 50.7 − 0.2
The bridge would then have to be designed for the governing load model (LM 71,
SW/0, or SW/1) increased by 3%.

5.4.2 Case with Dynamic Analysis


A simple case with 8 effective degrees of freedom will be solved by using modal
superposition. The same example was already analysed in Sections 5.1.4, 5.2.2,
and 5.2.3. It deals with a simply supported bridge with a span, L, of 45 m. The area
of the section is 7.96 m2 , the inertia, I y , is 7.71 m4 , and the distance from the
top face of the section to the centroid is 0.75 m. The elastic modulus of concrete
is taken as 36.2 GPa. Additionally, an upper-bound superimposed dead load
equivalent to G2,sup = 137.6 kN/m is present. So, the total mass of the structure is
M = (7.96 ⋅ 2.5 + 13.76) ⋅ 45 = 1514.7 ton. The bridge will be subjected to the passing
of 14 bogies (each of which will be modelled as a single concentrated load), spaced
D = 22 m apart, each weighing 340 kN, at a speed of 176.8 km/h, which corresponds
to the resonant speed. The assumed damping index is 1%.
Even though the system has 8 effective degrees of freedom (if horizontal move-
ments are not considered – see Figure 5.22) it can be solved as two systems with
4 degrees of freedom by using symmetry and antimetry with respect to the central
vertical axis as shown in Figure 5.23.

Pi(t) Pi(t) Pi(t) Pi(t)


Pi(t)

mL/8 = M/8 mL/4 = M/4 mL/4 = M/4 mL/4 = M/4 mL/8 = M/8

θ1(t) θ5(t)
y2(t), θ2(t) y4(t) , θ4(t)
y3(t), θ3(t)

l = L/4
R = 2.5 Pi(t) R = 2.5 Pi(t)

Figure 5.22 Simplified model of the structure.


170 5 Dynamic Behaviour of High-Speed Railway Bridges

Pi(t) Pi(t) Pi(t)/2 Pi(t) Pi(t) Pi(t)/2

mL/8 = M/8 mL/4 = M/ mL/8 = M/8 mL/8 = M/8 mL/4 = M/4 mL/8 = M/8
Bar 12 Bar 23 Bar 12 Bar 23
4
θ1(t) θ1(t) θ3(t)
y2(t), θ2(t) y2(t), θ2(t)
º
y (t)
3

l = L/4 l = L/4
R = 2.5 Pi(t) R = 2.5 Pi(t)
(a) (b)

Figure 5.23 Symmetric and antimetric structural models. (a) Symmetric vibration modes.
(b) Antimetric vibration modes.

The stiffness matrix of a bar is given in Eq. (5.40):


⎡ K𝟏𝟏 K𝟏2 ⎤
⎢⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⎥
⎢⎡ 12 6 ⎤ ⎡ 12 6 ⎤ ⎥
⎢⎢ 2 ⎢− l2 l ⎥ ⎥
⎢⎢ l l⎥
⎥ ⎢ 6 ⎥⎥
⎢⎢ 6
4 ⎥ ⎢− 2 ⎥ ⎥⎥
EI ⎢⎣ l ⎦ ⎣ l ⎦
Kbar = ⎢
l ⎢ K2𝟏 K22 ⎥ (5.40)
⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⎥
⎢ ⎥
⎢⎡− 12 − 6 ⎤ ⎡ 12

6 ⎤⎥
⎢⎢ l2 l⎥ ⎢ l2 l ⎥⎥
⎢⎢ 6 ⎥ ⎢ 6 ⎥⎥
⎢⎢ 2 ⎥ ⎢− 4 ⎥⎥
⎣⎣ l ⎦ ⎣ l ⎦⎦

The stiffness matrix of the structure is given in Eq. (5.41):


⎡ y1 𝜃1 y2 𝜃2 y3 𝜃3 ⎤
⎢ ⎥
⎢ 12 6 − 12 6 0 0 ⎥
⎢ l2 l l2 l ⎥
⎢ 6 6 ⎥
⎢ 4 − 2 0 0 ⎥
⎢ l l ⎥
⎡(K𝟏𝟏 )12 (K𝟏𝟐 )12 𝛀 ⎤ ⎢ 12 6 24 12 6 ⎥
⎢(K ) (K ) + (K ) EI ⎢−
(K𝟏𝟐 )23 ⎥ = − 0 − 2 ⎥ (5.41)
⎢ 𝟐𝟏 12 𝟐𝟐 12 𝟏𝟏 23 2 l l2 l ⎥
⎥ l ⎢ l l
⎣ 𝛀 (K𝟐𝟏 )23 (K𝟐𝟐 )23 ⎦ ⎢ 6 6 ⎥
⎢ 2 0 8 − 2 ⎥
⎢ l l ⎥
⎢ 12 6 12 6⎥
⎢ 0 0 − 2 −
l l l2
− ⎥
l
⎢ ⎥
⎢ 0 0
6
2 −
6
4 ⎥⎦
⎣ l l
5.4 Application Examples 171

For the symmetric case, the boundary conditions are that the deflection at node
1 and the rotation at node 3 are nil. To impose these conditions, all the items of
the row and column corresponding to these degrees of freedom are set to zero
except for the term on the diagonal which is set to 1 (see Eq. (5.42)). These rows and
columns can then be deleted from the matrix when determining the eigenvalues and
eigenforms.

⎡y1 𝜃1 y2 𝜃2 y3 𝜃3 ⎤
⎢ ⎥
⎢1 0 0 0 0 0⎥ ⎡ 𝜃1 y2 𝜃2 y3 ⎤
⎢ ⎥ ⎢ ⎥
⎢0 6 ⎢ 4 −6 2
4 − 2 0 0⎥ 0 ⎥
⎢ l ⎥ ⎢ l ⎥
⎢ 6 24 12 ⎥ ⎢ 6 24 12 ⎥
EI ⎢ 0 − 0 − 0 ⎥ → EI ⎢− 0 − 2⎥
l l2 l2 2 (5.42)
l ⎢ ⎥ l ⎢ l l l ⎥
⎢ 6 ⎥ ⎢ 6⎥
⎢0 2 0 8 −
l
0⎥ ⎢ 2 0 8 − ⎥
l
⎢ ⎥ ⎢ 12 6 12 ⎥⎥
⎢0 12 6 12
0 − 2 − 0 ⎥ ⎢ 0 − 2 −
⎢ l l l2 ⎥ ⎣ l l l2 ⎦
⎢ ⎥
⎣0 0 0 0 0 1⎦
For the antisymmetric case, the boundary conditions are that the deflection at
nodes 1 and 3 are nil (see Eq. (5.43)).
⎡y1 𝜃1 y2 𝜃2 y3 𝜃3 ⎤
⎢ ⎥
⎢1 0 0 0 0 0⎥ ⎡ 𝜃1 y2 𝜃2 𝜃3 ⎤
⎢ 6 ⎥ ⎢ 6 ⎥
⎢0 4 − 2 0 0⎥ ⎢ 4 − 2 0⎥
⎢ l ⎥ ⎢ l ⎥
EI ⎢ 6 24 6⎥ EI ⎢ 6 24 6⎥

l ⎢⎢
0 0 ⎥
l ⎢⎢ l ⎥
0 − − 0 (5.43)
l l2 l⎥ l2 l⎥
⎢0 2 0 8 0 2⎥ ⎢ 2 0 8 2⎥
⎢ ⎥ ⎢ ⎥
⎢0 0 0 0 1 0⎥ ⎢ 6 ⎥
⎢ ⎥ ⎣ 0 2 4⎦
6 l
⎢0 0 2 0 4 ⎥⎦
⎣ l
The mass will be concentrated at the nodes of the model. The concentrated mass
matrix is obtained by applying the following equations:
The translational mass for node i is given by:
(l1 + l2 )
mi = m (5.44)
2
The rotational mass for node i, is given by:
(l1 + l2 ) (l + l )
j i = 𝜌c I y + mSDL y2g 1 2 (5.45)
2 2
172 5 Dynamic Behaviour of High-Speed Railway Bridges

where:
l1 , l2 are the lengths of the two bars that converge on the node (one of which may be
zero if there is no bar)
m is the mass per length of the bridge including the self-weight and the superim-
posed dead load (ballast, track, barriers, communications)
𝜌c is the density of concrete
mSDL is the mass per length of the superimposed dead load (ballast, track, barriers,
communications)
yg is the distance from the centre of gravity of the superimposed dead load to the
centroid of the cross section.
The concentrated mass matrix can then be written as shown in Eq. (5.46):

⎡ y1 𝜃1 y2 𝜃2 y3 𝜃3 ⎤
⎢ l ⎥
⎢ m 0 0 0 0 0 ⎥
⎢ 2 ⎥
⎢⏟⏟ ⏟ ⎥
⎢ m1 ⎥
⎢ ( ) l ⎥
⎢ 0 𝜌c Iy + mSDL y2g 0 0 0 0 ⎥
⎢ 2 ⎥
⎢ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⎥
⎢ j1 ⎥
⎢ ⎥
⎢ 0 0 ml 0 0 0 ⎥
⎢ ⏟⏟ ⏟ ⎥
⎢ m2 ⎥
⎢ 0 ( ) ⎥
0 0 𝜌c Iy + mSDL yg l2
0 0
⎢ ⎥
⎢ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⎥
⎢ j2 ⎥
⎢ l ⎥
⎢ 0 0 0 0 m 0 ⎥
⎢ 2 ⎥
⎢ ⏟⏟⏟ ⎥
⎢ m3 ⎥
⎢ ( ) l ⎥
⎢ 0 0 0 0 0 𝜌c Iy + mSDL y2g ⎥
⎢ 2⎥
⎢ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⎥
⎣ j3 ⎦
(5.46)
Since the mass matrix is diagonal, its inverse, M−1 , will be the diagonal matrix
formed by the inverse values of the diagonal elements.
Now that matrices K and M−1 are known it is necessary to determine the eigenval-
ues of M−1 K, which will be equal to the square of the natural periods of the different
vibrations forms of the structure. The analysis for the symmetric case will provide
the eigenvalues of the symmetric eigenforms, while the analysis for the antimetric
case will provide the eigenvalues corresponding to the antimetric eigenforms.
Assuming that M is a diagonal matrix, multiplication by M−1 will just divide each
row of matrix K by mii .
5.4 Application Examples 173

Matrix M−1 K for the symmetric case can be written as follows:

1
⎡ 0 0 ⎤
0 ⎡ 4 −6 2 0 ⎤
⎢ j1 ⎥ ⎢ l ⎥
⎢ 1 ⎥ ⎢ 6 24 12 ⎥
⎢0 m 0 0 ⎥ ⎢− 0 − 2⎥
EI 2
M K =⎢
−𝟏 2 ⎥ ⎢ l l l ⎥
⏟⏟⏟ ⎢ 0 0 1
0 ⎥⎥ l ⎢⎢ 2 0 8 − ⎥
6⎥
𝛼A ⎢ j2 l
⎢ ⎥ ⎢ ⎥
⎢0 0 1 ⎥ ⎢ 12 6 12 ⎥
⎣ 0 0 − −
m3 ⎦ ⎣ l2 l l2 ⎦
⎡ 4 −
6 2
0 ⎤
⎢ j1 lj1 j1 ⎥
⎢ 6 24 12 ⎥
⎢− 0 − ⎥
EI ⎢ lm2 l2 m2 l2 m2 ⎥
l ⎢ 2 6 ⎥⎥
= (5.47)
8
⏟⏟⏟⎢ j 0 −
⎢ 2 j2 lj2 ⎥
𝛼
⎢ 12 6 12 ⎥
⎢ 0 − − ⎥
⎣ l m3 lm3 l m3 ⎦
2 2
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
A

It is a well-known property of Matrices that the eigenvalues of 𝛼A are equal to the


eigenvalues of A multiplied by 𝛼 so the EI/l factor can be kept outside the matrix
when deriving the eigenvalues.
For this example, the following numerical values result:

11.25
m1 = m3 = 33.66 ⋅ = 189.338 t
2
m2 = 2m1 = 378.675 t
11.25
j1 = j3 = (2.5 ⋅ 7.96 + 13.76 ⋅ 0.752 ) = 155.475 tm2
2
j2 = 2j1 = 310.95 tm2
EI 36.2 ⋅ 106 ⋅ 7.71
= = 2.481 ⋅ 107
l 11.25
M−𝟏 K = 2.481 ⋅ 107
⏟⏟⏟ ⏟⏞⏞⏞⏟⏞⏞⏞⏟
𝛼A 𝛼

⎡ 2.573 ⋅ 10−2 −3.430 ⋅ 10−3 1.286 ⋅ 10−2 0 ⎤


⎢ ⎥
⎢−1.408 ⋅ 10 −3
5.008 ⋅ 10 −4
0 −2.504 ⋅ 10 ⎥ −4
⋅⎢ ⎥ (5.48)
⎢ 6.432 ⋅ 10 2.573 ⋅ 10−2 −1.715 ⋅ 10−3 ⎥
−3
0
⎢ 0 −5.008 ⋅ 10−4 −2.817 ⋅ 10−3 5.008 ⋅ 10−4 ⎥⎦

⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
A
174 5 Dynamic Behaviour of High-Speed Railway Bridges

To solve for the eigenvalues of matrix A, it is necessary to obtain first the charac-
teristic equation. This is done in Equations (5.50) and (5.51).
| −1 |
| M K − 𝜆I| = 0
|⏟⏟⏟ |
A

|2.573 ⋅ 10−2 − 𝜆 −3.430 ⋅ 10−3 1.286 ⋅ 10−2 0 |


| |
| |
| −1.408 ⋅ 10 −3
5.008 ⋅ 10 −4
− 𝜆 0 −2.504 ⋅ 10 −4 |
| |
→| |
| 6.432 ⋅ 10−3 0 2.573 ⋅ 10 −2
𝜆 ⋅ 10 −3 |
| − −1.715 |
| |
| ⋅ −4
⋅ −3
⋅ −4
𝜆 |
| 0 −5.008 10 −2.817 10 5.008 10 − |
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
A
|5.008 ⋅ 10−4 − 𝜆 0 −2.504 ⋅ 10−4 ||
|
| |
⋅ (2.573 ⋅ 10−2 − 𝜆) || 0 2.573 ⋅ 10−2 − 𝜆 −1.715 ⋅ 10−3 ||
| |
| −5.008 ⋅ 10−4 −2.817 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆|
| |
|−3.430 ⋅ 10−3 1.286 ⋅ 10−2 0 |
| |
|
−3 |
|
+ 1.408 ⋅ 10 | 0 2.573 ⋅ 10 − 𝜆 −1.715 ⋅ 10 ||
−2 −3
| |
|−5.008 ⋅ 10−4 −2.817 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆|
| |
| −3.430 ⋅ 10−3 1.286 ⋅ 10−2 0 |
| |
|
−3 |
|
+ 6.432 ⋅ 10 |5.008 ⋅ 10−4 − 𝜆 0 −2.504 ⋅ 10−4 ||
| |
| −5.008 ⋅ 10−4 −2.817 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆|
| |
|2.573 ⋅ 10−2 − 𝜆 −1.715 ⋅ 10−3 |
| |
= (2.573 ⋅ 10−2 − 𝜆)(5.008 ⋅ 10−4 − 𝜆) | |
| −2.817 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆|
| |
| 0 −2.504 ⋅ 10−4 ||
|
+ (2.573 ⋅ 10−2 − 𝜆) ⋅ −5.008 ⋅ 10−4 | |
|2.573 ⋅ 10−2 − 𝜆 −1.715 ⋅ 10−3 |
| |
|2.573 ⋅ 10−2 − 𝜆 −1.715 ⋅ 10−3 |
| |
+ 1.408 ⋅ 10−3 ⋅ −3.430 ⋅ 10−3 | |
| −2.817 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆|
| |
| 1.286 ⋅ 10−2 0 |
| |
+ 1.408 ⋅ 10−3 ⋅ −5.008 ⋅ 10−4 | |
|2.573 ⋅ 10−2 − 𝜆 −1.715 ⋅ 10−3 |
| |
|5.008 ⋅ 10−4 − 𝜆 −2.504 ⋅ 10−4 |
| |
+ 6.432 ⋅ 10−3 ⋅ −1.286 ⋅ 10−2 | |
| −5.008 ⋅ 10−4 5.008 ⋅ 10−4 − 𝜆|
| |
| −3.430 ⋅ 10−3 0 |
| |
+ 6.432 ⋅ 10−3 ⋅ 2.817 ⋅ 10−3 | |=0 (5.49)
|5.008 ⋅ 10−4 − 𝜆 −2.504 ⋅ 10−4 |
| |
5.4 Application Examples 175

| −1 |
| M K − 𝜆I| = 0
|⏟⏟⏟ |
A

→ (2.573 ⋅ 10−2 − 𝜆)(5.008 ⋅ 10−4 − 𝜆)((2.573 ⋅ 10−2 − 𝜆)(5.008 ⋅ 10−4 − 𝜆) − 4.831 ⋅ 10−6 )

+ (2.573 ⋅ 10−2 − 𝜆)2 ⋅ −5.008 ⋅ 10−4 ⋅ 2.504 ⋅ 10−4


⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
−1.254 ⋅ 10−7

+ 1.408 ⋅ 10−3 ⋅ −3.430 ⋅ 10−3 ((2.573 ⋅ 10−2 − 𝜆)(5.008 ⋅ 10−4 − 𝜆) − 4.831 ⋅ 10−6 )
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
−4.829 ⋅ 10−6

+ 1.408 ⋅ 10−3 ⋅ −5.008 ⋅ 10−4 ⋅ 1.286 ⋅ 10−2 ⋅ −1.715 ⋅ 10−3


⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
−1.555 ⋅ 10−11

+ 6.432 ⋅ 10 ⋅ −1.286 ⋅ 10−2 ((5.008 ⋅ 10−4 − 𝜆)2 − 1.254 ⋅ 10−7 )


−3
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
−8.272 ⋅ 10−5

+ 6.432 ⋅ 10−3 ⋅ 2.817 ⋅ 10−3 ⋅ 8.589 ⋅ 10−7 = 0


⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
1.556⋅10−11

→ 𝜆4 + (−2.573 ⋅ 10−2 − 5.008 ⋅ 10−4 − 2.573 ⋅ 10−2 − 5.008 ⋅ 10−4 )𝜆3


⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
−5.246 ⋅ 10−2

⎛ ⎞
⎜ ⎟
⎜ 2.577⋅10 −5 6.620⋅10 −4 2.508⋅10 −7 ⎟
⎜ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⎟
⎜ ⎟
+ ⎜ 2 ⋅ 2.573 ⋅ 10−2 ⋅ 5.008 ⋅ 10−4 + (2.573 ⋅ 10−2 )2 + (5.008 ⋅ 10−4 )2 ⎟ 𝜆2
⎜ ⎟
⎜ −1.254 ⋅ 10−7 − 4.829 ⋅ 10−6 − 8.272 ⋅ 10−5 ⎟
⎜⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⎟
⎜ ⎟
⎝ 6.003 ⋅ 10−4 ⎠
⎛ ⎞
⎜ ⎟
⎜ −6.631⋅10−7 −1.291⋅10−8 ⎟
⎜ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⎟
⎜ ⎟
⎜ −2(2.573 ⋅ 10 ) ⋅ 5.008 ⋅ 10 −2 ⋅ 2.573 ⋅ 10 (5.008 ⋅ 10 )
−2 2 −4 −2 −4 2

⎜ ⎟
⎜ 1.243⋅10−7 2.419⋅10−9 6,53⋅10−9 ⎟
+ ⎜ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⎟ 𝜆
⎜ +2.573 ⋅ 10−2 ⋅ 4.831 ⋅ 10−6 + 5.008 ⋅ 10−4 ⋅ 4.831 ⋅ 10−6 + 2 ⋅ 1.254 ⋅ 10−7 ⋅ 2.573 ⋅ 10−2 ⎟
⎜ ⎟
⎜ +4.829 ⋅ 10−6 ⋅ (2.573 ⋅ 10−2 + 5.008 ⋅ 10−4 ) + 2 ⋅ 8.272 ⋅ 10−5 ⋅ 5.008 ⋅ 10−4 ⎟
⎜ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⎟
⎜ −8

⎜ 1.267 ⋅ 10−7 8.285⋅10 ⎟
⎜⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⎟
⎝ −3.413⋅10 −7 ⎠

1.660⋅10−10 −6.225⋅10−11
⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞
+(2.573 ⋅ 10−2 )2 (5.008 ⋅ 10−4 )2 −4.831 ⋅ 10−6 ⋅ 2.573 ⋅ 10−2 ⋅ 5.008 ⋅ 10−4
−8.302⋅10−11 −6.224⋅10−11 2.333⋅10−11
⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞
(−2.573 ⋅ 10−2 )2 ⋅ 1.254 ⋅ 10−7 −4.829 ⋅ 10−6 ⋅ 2.573 ⋅ 10−2 ⋅ 5.008 ⋅ 10−4 + 4.829 ⋅ 10−6 ⋅ 4.831 ⋅ 10−6
−2.075⋅10−11 1.037⋅10−11
⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞ ⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞
+1.555 ⋅ 10−11 −8.272 ⋅ 10−5 (5.008 ⋅ 10−4 )2 + 8.272 ⋅ 10−5 ⋅ 1.254 ⋅ 10−7 + 1.556 ⋅ 10−11
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
2.550⋅10−12

=0

→ 𝜆4 − 5.246 ⋅ 10−2 𝜆3 + 6.003 ⋅ 10−4 𝜆2 − 3.413 ⋅ 10−7 𝜆 + 2.550 ⋅ 10−12 = 0 (5.50)


176 5 Dynamic Behaviour of High-Speed Railway Bridges

8.00E-08 Symmetric system

6.00E-08 ω 32 = 419273

4.00E-08
Value of characteristic polynomial

ω 12 = 13769
2.00E-08 ω 12 = 197.2

0.00E+00
1.00E+00 1.00E+01 1.00E+02 1.00E+03 1.00E+04 1.00E+05 1.00E+06
–2.00E-08

–4.00E-08
ω 42 = 868317
–6.00E-08

–8.00E-08

–1.00E-07
λ = ω2

Figure 5.24 Characteristic polynomial and its solutions (symmetric system).

Figure 5.24 shows a plot of the characteristic equation where the 4 eigenvalues,
multiplied by EI/l are shown. These values correspond to the square of the first four
natural frequencies corresponding to symmetric vibration modes. The horizontal
scale is logarithmic. Figure 5.25 shows a detail of the first part of the curve where it
is easier to see the first two roots.

1.00E-10 Symmetric system (zoom)


Value of characteristic polynomial

5.00E-11

ω 12 = 197.2 ω 22 = 13769

0.00E+00
1.00E+00 1.00E+01 1.00E+02 1.00E+03 1.00E+04

–5.00E-11

–1.00E-10
λ = ω2

Figure 5.25 Characteristic polynomial and its solutions (symmetric system) – detail.
5.4 Application Examples 177

For the antimetrical case, an analogous procedure leads to the following


expressions:
⎡ 𝜃1 y2 𝜃2 𝜃3 ⎤
⎢ 4 6 2 ⎥
⎢ − 0 ⎥
⎢ 1j lj1 j1 ⎥
⎢ 6 24 6 ⎥
EI ⎢− 0
lm2 l2 m2 lm2 ⎥
l ⎢ ⎥
⎢ 2 0
8 2 ⎥
⎢ j2 j2 j2 ⎥
⎢ ⎥
⎢ 0 6 2 4 ⎥
⎣ lj3 j3 j3 ⎦

M−𝟏 K = 2.481 ⋅ 107


⏟⏟⏟ ⏟⏞⏞⏞⏟⏞⏞⏞⏟
𝛼A 𝛼

⎡ 2.573 ⋅ 10−2 −3.430 ⋅ 10−3 1.286 ⋅ 10−2 0 ⎤


⎢ −3 ⎥
⎢ −1.408 ⋅ 10 −3
5.008 ⋅ 10 −4
0 1.408 ⋅ 10 ⎥ (5.51)
⎢ 6.432 ⋅ 10−3 0 2.573 ⋅ 10−2 6.432 ⋅ 10−3 ⎥
⎢ 0 3.430 ⋅ 10−3 1.286 ⋅ 10−2 2.573 ⋅ 10−2 ⎥⎦

⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
A

The characteristic polynomial is determined, as above from:


| −1 |
| M K − 𝜆I| = 0 →
|⏟⏟⏟ |
A
𝜆 − 7.770 ⋅ 10−2 𝜆3 + 1.850 ⋅ 10−3 𝜆2 − 1.320 ⋅ 10−5 𝜆 + 1.600 ⋅ 10−9 = 0
4
(5.52)
Figure 5.26 shows a graph of the characteristic equation for the antimetric model,
showing also the four corresponding roots.

3.00E-08 Antimetric system

ω 32 = 647517
2.00E-08
Value of characteristic polynomial

1.00E-08 ω 12 = 320037
ω 12 = 3076

0.00E+00
1.00E+00 1.00E+01 1.00E+02 1.00E+03 1.00E+04 1.00E+05 1.00E+06

–1.00E-08

–2.00E-08

–3.00E-08 ω 42 = 957630
λ = ω2

Figure 5.26 Characteristic polynomial and its solutions (antimetric system).


178 5 Dynamic Behaviour of High-Speed Railway Bridges

For the symmetric model, the natural modes of vibration can then be derived from
the system of linear equations shown in Eq. (5.53), by substituting 𝜆 by 𝜔i 2 l/EI. In
this way the constant EI/l can be ignored.
( )
2 l
(M K − 𝜆i I)yi = 𝟎 ⇔ A − 𝜔i I yi = 𝟎
−𝟏
EI

⎤ ⎡𝜃1 ⎤ ⎡ ⎤
0
⎡2.573 ⋅ 10−2 − 𝜆 −3.430 ⋅ 10−3 1.286 ⋅ 10−2 0
⎢ ⎥ ⎢ ⎥ ⎢0⎥
⎢ −1.408 ⋅ 10
−3
5.008 ⋅ 10−4 − 𝜆 0 −2.504 ⋅ 10−4 ⎥ ⎢y2 ⎥ ⎢ ⎥
2.573 ⋅ 10 − 𝜆 −1.715 ⋅ 10−3 ⎥ ⎢𝜃2 ⎥ ⎢⎢ ⎥⎥
= 0
⎢ 6.432 ⋅ 10 −3
0 −2
⎢ 0
⎣ 0 −5.008 ⋅ 10−4 −2.817 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆⎥⎦ ⎢⎣y3 ⎥⎦ ⎢ ⎥
⎣0⎦
(5.53)

For the antimetrical model this can be done by the same procedure from the
expression of Eq. (5.54):

⎤ ⎡𝜃1 ⎤ ⎡ ⎤
0
⎡2.573 ⋅ 10−2 − 𝜆 −3.430 ⋅ 10−3 1.286 ⋅ 10−2 0
⎢ ⎥⎢ ⎥ ⎢0⎥
⎢ −1.408 ⋅ 10
−3
5.008 ⋅ 10−4 − 𝜆 0 1.408 ⋅ 10−3 ⎥ ⎢y2 ⎥ ⎢ ⎥
2.573 ⋅ 10−2 − 𝜆 6.432 ⋅ 10−3 ⎥ ⎢𝜃2 ⎥ ⎢⎢ ⎥⎥
= 0
⎢ 6.432 ⋅ 10−3 0
⎢ ⎥ ⎢ ⎥ 0
⎣ 0 3.430 ⋅ 10−3
1.286 ⋅ 10−2
2.573 ⋅ 10 − 𝜆⎦ ⎣y3 ⎦ ⎢ ⎥
−2
⎣0⎦
(5.54)

Since 𝜆 is an eigenvector of the matrix, the system of equations is not determined.


Therefore, the value of one of the eigenvectors components can be set to 1.00. For
instance, for the symmetric system, a value of y3 = 1.00 can be set. The system to be
solved would then be:

⎡2.573 ⋅ 10−2 − 𝜆 −3.430 ⋅ 10−3 1.286 ⋅ 10−2 ⎤ ⎡𝜃1 ⎤ ⎡ 0 ⎤


⎢ −1.408 ⋅ 10−3 5.008 ⋅ 10−4 − 𝜆 0 ⎥ ⎢ y2 = 2.504 ⋅ 10−4 ⎥ (5.55)
⎥ ⎢
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣ 6.432 ⋅ 10−3 0 2.573 ⋅ 10−2 − 𝜆⎦ ⎣𝜃2 ⎦ ⎣−1.715 ⋅ 10−3 ⎦

The corresponding eigenvectors for the symmetric structure are the following:

⎡ 𝜔2i ⎤ ⎡ 197.2 13769 419273 868317⎤


⎢𝜃 ⎥ ⎢0.0697 −0.151 8.07 −17.5 ⎥
⎢ 1,i ⎥ ⎢ ⎥
⎢y2,i ⎥ = ⎢ 0.707 −0.707 −0.707 0.707 ⎥ (5.56)
⎢𝜃2,i ⎥ ⎢0.0493 0.107 −5.71 −12.4 ⎥
⎢ ⎥ ⎢ ⎥
⎣y3,i ⎦ ⎣ 1 1 1 1 ⎦

For the antimetric structure, the eigenvectors are:

⎡ 𝜔2i ⎤ ⎡ 3076 320037 647517 957630⎤


⎢𝜃 ⎥ ⎢ −1 1 −1 1 ⎥
⎢ 1,i ⎥ ⎢ ⎥
⎢y2,i ⎥ = ⎢−7.47 0 −0.11 0 ⎥ (5.57)
⎢𝜃2,i ⎥ ⎢ 0 −1 0 1 ⎥
⎢ ⎥ ⎢ ⎥
⎣𝜃3,i ⎦ ⎣ 1 1 1 1 ⎦
5.4 Application Examples 179

The full eigenvector matrix will therefore be:

⎡ 𝜔2i ⎤ ⎡ 197.2 3072 13769 320037 419273 647517 868317 957630⎤


⎢𝜃 ⎥ ⎢ 0.0697 −1 0.151 1 8.07 −1 −17.5 1 ⎥
⎢ 1,i ⎥ ⎢ ⎥
⎢y2,i ⎥ ⎢ 0.707 −7.47 −0.707 0 −0.707 0.11 0.707 0 ⎥
⎢𝜃 ⎥ ⎢ 0.0493 0 0.107 −1 5.71 0 −12.4 1 ⎥
⎢ 2,i ⎥ ⎢ ⎥
⎢y3,i ⎥ = ⎢ 1 0 1 0 1 0 1 0 ⎥
⎢𝜃3,i ⎥ ⎢ 0 1 0 1 0 1 0 1 ⎥
⎢ ⎥ ⎢ ⎥
y
⎢ 4,i ⎥ ⎢ 0.707 7.47 −0.707 0 −0.707 −0.11 0.707 0 ⎥
⎢𝜃4,i ⎥ ⎢−0.0493 0 −0.107 −1 −5.71 0 12.4 0 ⎥
⎢ ⎥ ⎢ ⎥
𝜃
⎣ 5,i ⎦ ⎣ −0.0697 −1 −0.151 1 −8.07 −1 −17.5 1 ⎦
(5.58)
Figure 5.27 shows the shape of the first 3 eigenforms.
The natural periods of vibration can be determined as shown in Eq. (5.59)
⎧0.447 s

⎪0.113 s
⎪5.35 ⋅ 10−2 s
√ ⎪
1 ⎪1.11 ⋅ 10−2 s
Ti = 2π =⎨ (5.59)
𝜔2
⎪9.70 ⋅ 10
−3
s
⎪7.81 ⋅ 10−3 s

⎪6.74 ⋅ 10
−3
s
⎪6.42 ⋅ 10−3 s

The first natural period practically coincides, as is to be expected, with the value
obtained from the equivalent SDOF estimated in Section 5.1.4 (0.44 s).

2.00

1.50
Eigen vector shape

1.00

0.50

0.00
0.00 5.00 10.00 15.00 20.00 25.00 30.00 35.00 40.00 45.00
–0.50

–1.00

–1.50

–2.00
x (m)
First vibration mode Second vibration mode Third vibration mode

Figure 5.27 Shape of 3 eigenvectors.


180 5 Dynamic Behaviour of High-Speed Railway Bridges

400 Force history applied on nodes 2, 3, and 4 for the first vibration mode

350

300

250
ϕj1Fj (kN)

200

150

100

50

0
0.00 0.20 0.40 0.60 0.80 1.00
t (s)

Node 2 Node 3 Node 4

Figure 5.28 Load history for nodes 2, 3, and 4 as a single axle weighing 340 kN crosses
the bridge, weighted by the eigenform – First vibration mode.

The next step is to solve, for each eigenfrequency and eigenmode, i, the equation
of motion for a single load circulating on the bridge, and then obtain the response
to the train by summing up the effect of the consecutive axles.

Ci∗ Ki∗ ∑
n
𝜙ji
q̈ i + ̇
∗ qi + ∗ qi = Fj (t)jy (5.60)
Mi Mi j=1
Mi∗

For each node, the effect of the passing single load is modelled with a triangular
load. The load history of the three nodes with vertical movement of the model is
represented in Figure 5.28. The analysis can be further simplified by considering the
load applied on each node separately and then adding the deflections obtained from
the three equations:

Ci∗ Ki∗ 𝜙2i


q̈ i + ̇
∗ qi + qi = F (t)
Mi Mi∗ Mi∗ 2
Ci∗ Ki∗ 𝜙3i
q̈ i + q̇ i + qi = F (t)
Mi∗ Mi∗ Mi∗ 3
Ci∗ Ki∗ 𝜙4i
q̈ i + q̇ i + qi = F (t) (5.61)
Mi∗ Mi∗ Mi∗ 4

The values of Ci∗ , Mi∗ , and Ki∗ can be determined by forward and backward
multiplication of the Damping, Mass, and Stiffness matrices by the corresponding
eigenvalue. Alternatively, and more easily, Ci∗ can be determined as:

Ci∗ = 2𝜉i 𝜔i Mi∗ (5.62)


5.4 Application Examples 181

Table 5.5 Parameters to be used for first bogie and the first mode of
vibration in the formulations of Appendix A.

Node #, j t0 t 0 + 𝚫t/2 t 0 + 𝚫t 𝝓j,1

2 0.00 0.229 0.458 0.707


3 0.229 0.458 0.687 1.00
4 0.458 0.687 0.916 0.707

The resulting values of coefficients Ci∗ (assuming all modes have the same 2%
damping index), Mi∗ , and Ki∗ for the symmetric model are given in Eq. (5.63):
⎡ 212.94 ⎤ ⎡ 380.19 ⎤ ⎡ 7.454 ⋅ 104 ⎤
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
1.809 ⋅ 103 ⎥ 385.76 ⎥ 5.305 ⋅ 106 ⎥
C∗ = ⎢ M∗ = ⎢ K∗ = ⎢ (5.63)
⎢5.350 ⋅ 105 ⎥ ⎢2.064 ⋅ 104 ⎥ ⎢ 8.669 ⋅ 109 ⎥
⎢3.553 ⋅ 106 ⎥ ⎢9.538 ⋅ 104 ⎥ ⎢8.274 ⋅ 1010 ⎥
⎣ ⎦ ⎣ ⎦ ⎣ ⎦
The solution to the dynamic equation of a single degree of freedom system sub-
jected to a triangular load is solved in [3] and presented in Appendix A. However,
these formulations do not account for damping. Damping is important when there
is a series of axles crossing the bridge since it will significantly reduce resonance
effects. Therefore, a generalisation of these expressions is included in Appendix A.
The differential equations shown in Eq. (5.61) can be solved using the solution
detailed in Appendix A, applied to the values for t0 , Δt/2 and Δt shown in Table 5.5
for the first vibration mode:
To determine the maximum deflection at mid-span (which corresponds to the
third degree of freedom) the following expression will be applied:

n
y3 = 𝜙3,i qi (5.64)
i=1

Observation of Eq. (5.58) leads to the conclusion that only modes 1, 3, 5, and 7 will
contribute to the vertical motion of the central node since the coefficients 𝜙i,j corre-
sponding to the other modes for the row corresponding to y3,i are nil. Additionally,
because the deflection of a given mode is inversely proportional to natural frequency
of the damped system, 𝜔d , because the damping of the system occurs faster when
the natural frequency is high, and because the frequency increases quite rapidly
(factor
√ of 8.36 – theoretically this should be 9 – from the first mode to the third:
13769
∼ 8.36), the contribution of the higher modes will be neglectable (see below).
197.2
Figure 5.29 shows the deflection in the structure when a 340 kN load crosses the
bridge at 177 km/h considering only the effects of vibration mode 1. The result is
the sum of the effect of the triangular pulse acting on nodes 2, 3, and 4 multiplied
by the modal factor 𝜙j,1 , where j = 2, 3, 4. In this case 𝜙j,1 = [0.707;1.00;0.707]. By
comparison with Figure 5.13, it is apparent that the same results are obtained with
the matrix formulation as with assumed sinusoidal eigenforms. However, the matrix
formulation is more general as it can be applied to any structure without previous
knowledge of the eigenform shapes.
182 5 Dynamic Behaviour of High-Speed Railway Bridges

340 kN load travelling over a 45 m simply supported span at 177 km/h


Modal analysis - Decomposition of contribution of load acting on each node
4.00

3.00

2.00

1.00
y (mm)

0.00
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00 4.50 5.00
–1.00

–2.00

–3.00
Time (s)

Sum of effects on nodes 2, 3, and 4 Axis 1 acting on node 2 · 0.707


Axis 1 acting on node 3 · 1.00 Axis 1 acting on node 4 · 0.707

Figure 5.29 Effect of single load travelling over bridge at 177 km/h – first mode only.

340 kN load travelling over a 45 m simply supported span at 177 km/h


Modes 1 and 3 (different scales)
3.00 0.06

2.50 0.05

2.00 0.04

1.50 0.03 ymode3 (mm)


ymode1 (mm)

1.00 0.02

0.50 0.01

0.00 0
0.00 0.50 1.00 1.50 2.00
–0.50 –0.01

–1.00 –0.02

–1.50 –0.03
Time (s)

Mode #1 Mode #3

Figure 5.30 Comparison of the contribution of the first and third modes.

Figure 5.30 shows a comparison of the contribution of the response of the structure
to the passing of one bogie of the train of modes 1 and 3. Note that the contribution
of mode 3 is plotted using the secondary vertical axis and therefore both contribu-
tions are not represented on the same scale. The maximum deflection from mode
1 is close to 2.7 mm while the maximum deflection from mode 3 is only 0.033 mm,
that is around 1% of the previous figure. It can be concluded as in Section 5.2.2 that
References 183

Superposition of 7 and 14 bogies compared with the effect of one


15.00
13.00
11.00
9.00
7.00
5.00
3.00
y (mm)

1.00
–1.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.00
–3.00
–5.00
–7.00
–9.00
–11.00
–13.00
Time (s)
1 Bogie 7 Bogies 14 Bogies

Figure 5.31 Variation of the maximum deflection with time due to the passage of several
bogies.

sufficient accuracy can be obtained by considering only the first vibration mode for
the case of this simply supported beam.
By superimposing the effect of several bogies circulating on the bridge with a time
offset of D/v = 22/177 ⋅ 3.6 = 0.447 s, the effect of the train load can be determined.
Figure 5.31 shows the results comparing the effect of a single bogie to that of 7, to
that of 14. The results are nearly identical to those obtained with assumed sinusoidal
eigenshape functions (see Figure 5.14). The maximum deflection considering 14
bogies is 13.8 mm, while the maximum static deflection which occurs when two
bogies are on the bridge (since D = 22 m and L = 45 m) is 3.27 mm. This provides a
DLF of 4.2.

References

1 CEN (2003), EN 1991-2 Eurocode 1 – Actions on structures – Part 2: Traffic loads


on bridges.
2 UIC – Union Internationale des Chemins de Fer. (1979). Charges a prendre en
considerations dans le calcul des ponts-rail. Code UIC 776-1R.
3 Biggs, J. (1964). Introduction to Structural Dynamics. New York: McGraw-Hill
Book Company.
4 Domínguez Barbero, J. (2001. Dinámica de puentes de ferrocarril para alta
velocidad: métodos de cálculo y estudio de la resonancia. PhD Thesis. Madrid:
Technical University of Madrid.
5 Ling, L., Xiao, X.-b., Xiong, J.-y. et al. (2014). A three-dimensional model for
coupling dynamics analysis of high speed train-track system. Journal of Zhejiang
University-SCIENCE A (Applied Physics & Engineering).
185

Longitudinal Track–Structure Interaction


Manuel Cuadrado and Alejandro Pérez-Caldentey

6.1 Introduction
This chapter deals with the interaction between track and structure. Due to move-
ments related to loads (vertical loads and traction and braking forces), tempera-
ture variations and the rheological behaviour of concrete, differential displacements
occur between the track and the structure. These differential movements generate
additional stresses in the rails, with respect to the situation of the rail on the ground,
which could lead to failure of the rails, and can affect the safety and serviceability
of the train circulation. This chapter describes how to analyse these effects, which
loads need to be considered, how they are combined, and which verifications need
to be made.
If the rail stress verifications are not satisfied, then a track joint has to be intro-
duced. This is a measure that ought to be avoided as much as possible due to the
maintenance costs associated with track joints. However, in some cases they are
unavoidable. In such cases, the track joints need to be designed (allowed displace-
ment) and regulated. Installation and maintenance of track joints are also discussed.
Finally a series of considerations are given depending on the longitudinal layout of
the bridge with support from a series of design examples which are fully developed.
For more information cf. [1].

6.2 Problem Statement


To improve passenger comfort and reduce maintenance costs, modern railways use
a Continuous Welded Rail (CWR), which has the bare minimum number of joints.
Due to temperature at the ends of the rail a breathing length occurs, for a length
of about 150 m, where there is relative movement between rail and the ground and
where friction forces develop. After a length of about 150 m friction is enough to
stop any relative movement. As this point the stress in the rail becomes constant
and equal to the differential temperature of the rail with respect to the ambient tem-
perature times the modulus of elasticity of steel times the dilatation coefficient of rail

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
186 6 Longitudinal Track–Structure Interaction

Rail stresses Breathing


Friction forces
length

σs = EsαΔT

Jointless length

Track

Figure 6.1 Stresses in Continuous Welded Rail (CWR) placed ground between rail joints.

steel (see Figure 6.1). The stress in the rail due to the temperature difference with
respect to the mean ambient value will be given by Eq. (6.1). For a 50 ∘ C tempera-
ture difference the stresses in the rail are about 105 MPa. The track is designed to
withstand these stresses, residual stresses due to manufacturing process and those
coming from traffic loads, plus a limited overstress.

𝜎s = Es 𝛼ΔT (6.1)

When the railtrack crosses a bridge, this analysis is complicated by the relative
movements between track and structure. These relative movements will increase the
stresses in the rails and can lead to deconsolidation of ballast, in the case of ballasted
track, or excessive forces on other track components, such as fasteners, in the case of
ballastless track. The track–structure interaction analysis is carried out to verify that
the stresses remain within admissible ranges and that other effects due to relative
displacements are controlled.
The track and deck of a bridge are connected by a series of elements whose
behaviour is certainly non-linear (see Figure 6.2) and in reality very complex. For a
ballasted track, the rails are connected to the sleeper by a fastening systems which
includes clips, pads, tie plates, and bolts. The sleepers are embedded on a bed of
ballast with 30–70 cm of ballast below the bottom face of the sleeper.
The connection between deck and rail is strained for a number of reasons. First of
all the track and the deck of the bridge will be subjected to different imposed strains.
Due to temperature, while the rail will not move if there are no track joints, the
bridge deck will expand and contract due to temperature variations (which can reach
around 35 ∘ C). But even if there are track joints, because of the much greater thermal
conductivity of steel with respect to concrete, the rails will be subjected to larger
temperature variations (around 50 ∘ C) and therefore larger strain variations than the
deck, which will also be at least partially protected from direct sunlight by the track.
Additionally, the concrete deck will be subjected to strains due to the rheological
behaviour of concrete (creep and shrinkage) which will be absent from the rails.
Vertical traffic loads cause rotations in the deck which will increase relative dis-
placement between the deck top surface and the track, as the deck will rotate with
respect to its centroid or with respect to its bearings if the longitudinal displacement
of the bearing is fixed. Traction and braking loads will be applied on the rails and will
6.2 Problem Statement 187

7.00
2.80 4.20
1.30 1.50 1.85 2.35
Rail
Sleeper

Ballast
0.20
1.76

Figure 6.2 Half section of a high-speed railway bridge showing a ballasted track and the
deck.
be transmitted to the deck through the connection between rail and sleeper and then
will be transmitted horizontally by friction within the ballast to the bridge deck. The
bridge deck will then move depending on the stiffness of the piers and the presence
or absence of structural joints.
As a result of all these actions, overstresses will develop in the rails with respect
to a situation where the track is laid on the ground which does not move. If the
overstresses are too high there is a danger that the rails will fail, either in tension
or in compression by buckling, leading a catastrophic accident as a train crosses the
bridge.
Another issue that can cause significant problems is the relative displacements
between decks at structural joints and between abutments and the deck ends. Here
longitudinal and vertical relative displacements need to be limited. If these relative
displacements are too large some undesirable phenomena can occur. In the case of
ballasted tracks there is a risk of ballast deconsolidation that can lead to both safety
issues as well as to a deficient quality of the train ride. In the case of ballastless track
relative displacements can produce excessive efforts on the rail or fasteners, leading
to fatigue problems for these elements or problems for rail pads (that can get loose
if clamping forces are reduced).
If it is not possible to comply with the given limits to rail overstresses track joints
have to be introduced (see Figure 6.3). This measure, however, is to be avoided when-
ever possible due to the maintenance problems which track joints entail.
The importance of the effects of track–structure interaction will depend on the
structural configuration (expansion length, presence of joints in the deck), the
188 6 Longitudinal Track–Structure Interaction

Figure 6.3 Examples of track joints.

mechanical properties of the deck (vertical flexibility, rotation at the ends, distance
between the centroid and the top section, distance between centroid and bearings,
longitudinal stiffness of the bearing + pier system) as well as on the type of track
(ballasted or ballastless track), the presence of rail expansion devices, and on the
properties of the track (axial stiffness of the rails, resistance of the rails and sleepers
to longitudinal movements within the ballast, for ballasted tracks, or resistance of
the rail fastenings, for unballasted tracks).

6.3 Model for Analysis


6.3.1 General Considerations
Figure 6.4 shows a scheme of a model that can be used to model the track–structure
interaction. Different horizontal levels are considered for modelling:
– centroid of the rails
– top face of the deck
– centroid of the deck
– bottom of the deck
– level of foundations.
Nonlinear springs
Representing horizontal
Track stiffness
Top of deck
More than 200 M Centroid of rail
Centroid of deck More than 200 m

Springs representing Bottom of deck


Bearing stiffness

Pier

Figure 6.4 Possible model for the analysis of track–structure interaction (as suggested in
Figure 3.3 of reference [2]).
6.3 Model for Analysis 189

6.3.1.1 Rails
The rails are connected to the top of the deck by a rigid element in the vertical direc-
tion and by non-linear springs in the horizontal direction. These springs model the
response of the sleeper and ballast to horizontal loads (or the force-slip behaviour
for unballasted tracks). A very important feature of the model is that a significant
stretch of track outside the bridge needs also to be modelled in order to obtain real-
istic boundary conditions (fixity between track and ground outside the structure).
This length is recommended to be at least equal to 200 m, and, in any case, larger
than the breathing length. Additionally, [3] recommends that the distance between
nodes be limited to a maximum of 2.00 m.
The rails should be modelled as a beam element. All the rails on the structure’s
deck can be grouped into a single element having the sum of the area and inertia of
all the actual rails (except in the case of some exceptional geometries, as it will be
shown in Sections 6.7.6.2 and 6.7.6.3).

6.3.1.2 Deck
The deck is also modelled using beam elements with the area and inertia of the deck.
The top of the deck is connected by means of stiff vertical members to the centroid
of the deck. This allows to simulate the effects of rotations of the deck on the top of
deck, in terms of strains.
When modelling a series of simply supported precast spans with a continuous top
slab over several spans, a double-hinged element should be placed between spans
to model the longitudinal continuity of the deck at the level of the top slab, while
allowing deck rotations.

6.3.1.3 Interaction Between Rails and Track Base


According to current practice [3, 4], the connection between rails and deck is
modelled using elasto-plastic springs. For a ballasted track, the maximum load is
reached when the resistance of the ballast subjected to a horizontal load applied
by the sleepers is overpassed. For an unballasted track, it is reached when the
connection between rail and foundation slab slips. This resistance is given by length
of track, since it will increase as the number of sleepers or connections increases.
Figure 6.5 shows the force per meter of track (k) vs relative displacement between
track and the top face of the deck (u). The yield displacement (u0 ) depends on
whether the track is a ballasted track or an unballasted track. The maximum force
per meter of track depends on whether or not the track is loaded. This is because the
frictional resistance is enhanced by pressure. The elasto-plastic model is, of course,
a major simplification of a very complex problem. EN 1991-2 [4] does not provide
recommended values for the springs. Instead it labels these values as Nationally
Determined Parameters (NDP) to be adopted by the different countries. Table 6.1
shows the values recommended by UIC [3].

6.3.1.4 Bearings
At each support (abutments or piers) the centroid of the deck is connected to the
bottom of the deck where the structural bearings are. This point is connected to the
190 6 Longitudinal Track–Structure Interaction

70
Loaded
60

50 Ballasted
k (kN/m)

40
Unloaded
30

20

10 Without
ballast
0
0 2 4 6 8 10 12
u (mm)

Ballasted track: unloaded Ballasted track: loaded


No ballast: unloaded No ballast: loaded

Figure 6.5 Force per length vs longitudinal displacement.

Table 6.1 Values of track resistance, k, and yielding displacement, u0 , recommended by


UIC [5].

Unloaded Loaded

u0 (mm) k u (kN/m) k l (kN/m)

Ballasted Normal maintenance 2 12 60


Careful maintenance 2 12 60

No ballast, conventional fasteners 0.5 40 60

top of the abutment or pier by a vertical and a horizontal spring, which are meant
to model the stiffness of the deck–support connection. These can be linear springs if
the supports are neoprene bearings or non-linear springs if the supports are sliding
bearings. They can also be rigid if there is a monolithic connection

6.3.1.5 Columns
The columns are modelled as beam elements with the stiffness provided by the pier
cross sections. In normal practice, the uncracked pier stiffness is considered when
modelling the track–structure infrastructure.

6.3.1.6 Foundations
The model can be further complicated, by considering the rotational and transla-
tional flexibility of the foundations. This can be done by adding linear or non-linear
springs at the level of the foundations. However, given the highly non-linear
6.4 Actions 191

behaviour and the large uncertainties involved in the characterisation of the soil,
such refinements may be more of a theoretical exercise than a reflection of the
structure’s actual behaviour and will not necessarily lead to greater reliability.

6.4 Actions
6.4.1 Temperature Variations
It is necessary to consider the two components of thermal variation:
– Uniform variation. The variation of temperature in the rail can be taken as ±50 ∘ C.
The uniform variation of temperature of the deck can be determined according to
EN 1991-1-5 [6]. EN 1991-2 allows to consider a simplified value of ±35 ∘ C
– Temperature gradient in the deck, which can also be determined according to EN
1991-1-5. This gradient will induce rotations which will produce displacements at
the level of the top deck surface.
The value of k to be used to estimate the effects of temperature should be that of
the unloaded track (ku ), since the rate of application of this load will be much slower
than the transit of a train through the bridge.

6.4.1.1 Case Without Track Joint


The effect of the temperature of the rail depends on whether or not there is a rail
joint. If there is not, then the variation of temperature of the rail can be ignored,
as the track will not move since it is fully anchored by friction forces. Comparison
will be made in terms of increase of stresses with respect to the situation outside
the bridge, where the rail can be subjected to a stress of 105 MPa as demonstrated
in Section 6.2. So only the temperature variation and gradient in the deck will be
considered.

6.4.1.2 Case with Track Joint


If there is a joint in the track then the analysis will be made in terms of total stresses.
In this case, obviously the stress in the rail at the joint will be nil. Also the rails will
move and these movements will also affect the reactions and displacements of the
bridge.

6.4.2 Traction and Braking Forces


The traction and braking forces should be applied accounting for the permitted cir-
culation direction of each track. Therefore, in a bridge with two tracks with opposing
circulation directions the most unfavourable situation is to consider that a train is
braking on one track while a train is initiating movement on the other.
EN 1991-2 defines, for all loading models except SW/2, the braking force as a load
of 20 kN/m times the length over which the force is applied. The maximum value
for the braking force is 6000 kN, which effectively provides a maximum application
192 6 Longitudinal Track–Structure Interaction

length of 300 m. For SW/2 loading, the value is increased to 35 kN/m. However, in
this case the loaded length is limited to 57 m only.
Regarding traction, and this time for all load models, including SW/2, the force
is defined as 33 kN/m times the length of application, but not greater than 1000 kN,
which effectively limits the length of application to 30 m.
In this case, the value of k should depend on whether or not a given zone of the
structure is considered to be loaded.

6.4.3 Vertical Loads


The effect of vertical loads is twofold:
– On one hand, the vertical loads produce rotations which generate displacements
at the top level of the deck.
– On the other hand, the presence of the vertical load increases the maximum load
that can be taken by the springs that model the stiffness of the connection between
rail and deck, thereby increasing the maximum stress in the rails.
The relevant vertical load to be considered for the study of the track–structure
interaction is LM 71 and if required SW/0 and SW/2. According to EN 1991-2,6.5.4.3,
dynamic effects associated with the vertical load maybe neglected.
The effect of vertical loads is related to the maximum rotations at deck expan-
sion joints. The maximum rotations can be maximised by loading alternate spans.
This hypothesis is no doubt conservative. To consider more realistic situations, how-
ever, involves the consideration of a potentially very large number of non-linear load
combinations.
Regarding the value of k, as with the traction and braking forces, it should depend
on whether or not a given zone of the structure is considered to be loaded. To max-
imise the effects of vertical loads, one of the accesses to the bridge should be con-
sidered loaded. It does not make sense to also consider the other access as loaded
because the deck will be fully deflected as the train starts exiting from the bridge
and no increase in these deflections will occur as the train circulates also over the
bridge exit and therefore there will be no change in the corresponding forces and
displacements as k increases from ku to kl in the exit access due to the passing train.

6.4.4 Creep and Shrinkage


Creep and shrinkage will produce a shortening of the bridge deck, which will
increase the stresses in the rail. While in prestressed bridges, shrinkage will produce
mostly only axial displacements, creep will also produce rotations, which should
be accounted for. The part of the creep and shrinkage that should be considered
is only that part which occurs after the installation of the rails. Since a significant
time lapses between the end of construction and the rail installation, and since the
rate development of creep and shrinkage is decreasing through time, a significant
reduction can be achieved by not considering the creep and shrinkage that occurred
before the installation of the rail [7].
6.4 Actions 193

100

90
Equivalent temperature variation (°C)

80

70 45 °C
60

50

40 92 °C

30
47 °C
20
2 Years
10

0
1 10 100 1 000 10 000 100 000
Time (days)

Shrinkage Creep Creep + Shrinkage Series1 Series2

Figure 6.6 Example of creep and shrinkage strain occurring before the installation of the
rail at t = 2 years.

Figure 6.6 shows an example. A deck of concrete class C35/45 with a notional
depth equal to 500 mm, in an environment with a relative humidity of 50% with a
mean axial compression due to prestressing equal to 6 MPa, would be subjected to an
equivalent temperature load of 90 ∘ C. If the creep and shrinkage happening in the
first two years is discounted, the equivalent temperature variation drops to 45 ∘ C,
that is, half the total value.
Additionally, in ballasted tracks, track maintenance operations to keep the track
layout constant involve local lifting of the track. Such operations liberate a part of
the stresses due to creep and shrinkage, so that considering the total value of creep
and shrinkage deformations may be over-conservative, even after discounting the
deformation occurring before fixing of the rails. However for an unballasted track,
no further reductions are possible.
For shrinkage and creep, the value of k to be considered is that of the unloaded
track (ku ) since, as happens with temperature variations, these actions develop
slowly with time and are not affected by the momentary increase in resistance
afforded by a passing train.

6.4.5 Combination of Actions


Due to the non-linear behaviour of the connection between rail and deck, the prob-
lem cannot, in principle, be solved by linear superposition. However, since a full
non-linear analysis is too complicated, as they could need step-by-step computations
[8–11] (see Example 2 in Section 6.7.1.2), the following simplified and conservative
procedure is proposed:
194 6 Longitudinal Track–Structure Interaction

Table 6.2 Load combinations to be considered for track–structure interaction.

Envelope Envelope
Hypothesis Hypothesis (Alternate spans) (Positive + Negative)
Temperature Temp decrease + Traction
increase + negative positive gradient + Vertical and
gradient shrinkage + creep loads braking

Load 1 2 3 4
Comb 1 X X
Comb 2 X X
Comb 3 X X X
Comb 4 X X X

1. Determine the effects of a temperature increase, coupled with a negative gradient,


using ku (track unloaded)
2. Determine the combined effects of a temperature decrease, coupled with a posi-
tive gradient, plus creep and shrinkage, using ku (track unloaded)
3. Consider loading of alternate spans with vertical loads, as well as the loading
of one of the approach spans. In loaded spans use kl (track loaded). This will
provide 2 load scenarios as there are two possibilities of loading alternate spans.
Consider loading on all the tracks simultaneously. Track classification factors (𝛼)
should be considered for the vertical loads, but dynamic load factors (Φ) can be
neglected.
4. Apply maximum braking and traction forces, simultaneously (except if the struc-
ture has one track only. In this case apply braking forces only). Use kl in the length
assumed to be braking or starting. Traction and braking forces are not affected by
Φ but do have to be multiplied by 𝛼.

Consider the combinations of Table 6.2.

6.5 Verifications
6.5.1 Verifications in Terms of Stresses
EN 1991-2:2003 provides limits to the rail overstresses which are valid when the
following conditions are fulfilled:

– UIC 60 rail with a minimum strength of 900 MPa


– Track radius of at least 1500 m
– In case of ballasted tracks:
– Heavy concrete sleepers separated no more than 650 mm apart
– Minimum ballast depth under the sleeper of 300 mm.
6.5 Verifications 195

In a bridge without a track joint the maximum variation of stresses in the rails
should be limited to 72 MPa in compression and 92 MPa in tension. These are over-
stresses, since the continuous welded rail (CWB) would be subjected to 105 MPa (see
Section 6.2) without the presence of a bridge.
If a track joint is present, the verification should be made in terms of total stresses
in the rail. In this case, the limits would be 72 + 105 = 177 MPa in compression and
92 + 105 = 197 MPa in tension.

6.5.2 Verifications in Terms of Displacements


Due to the effects of traction and breaking only, the relative displacement between
two decks at an expansion joint of the deck or between the abutment and the end of
the deck (𝛿 B ) should be (see Figure 6.7):

– Less than 5 mm if the rail is a continuous welded rail or if there is a track expansion
device at one end of the deck
– Less than 30 mm if there are track expansion devices at both ends of the deck

Due to vertical loads only, the horizontal displacement (𝛿 H ) (see Figure 6.8) of the
deck at the end of the deck should be limited to:

– 8 mm if the deck rotation is determined accounting for the stiffness of the track
– 10 mm if the deck rotation is determined accounting only for the stiffness of the
deck.

This verification will be influenced by the typology of the bearing, since rotations
will produce larger displacements at the top of the deck if the bearings are fixed than
if the bearings are free to move longitudinally, as shown in Figure 6.8.

δB
δf
δi
δB δi
δf
δi

δB = ∣δf – δi∣

δB = ∣δf – δi∣
(a)

(b)

Figure 6.7 (a) Displacement between decks at expansion joint and (b) between deck and
abutment at bridge end.
196 6 Longitudinal Track–Structure Interaction

δH δH

(a) (b)

Figure 6.8 Horizontal displacement of deck due to vertical loads (a) fixed bearing;
(b) sliding bearing.

δv

δv

(a) (b)

Figure 6.9 (a) Vertical relative displacement between two deck or (b) between deck and
abutment due to longitudinal displacement coupled with longitudinal slope.

Due to vertical loads and temperature variations, the relative vertical displacement
between the end of two decks at an intermediate deck joint, or between the end span
and the abutment (see Figure 6.9) should be limited to:
– 3 mm if the maximum line speed at the site is less than or equal to 160 km/h
– 2 mm if the maximum line speed at the site is more than 160 km/h.
This discontinuity, in this case, is caused by the longitudinal slope of the deck, and
will, therefore, be sensitive to this variable.

6.5.3 Criteria for Placing a Track Joint


Track joints must be placed when the stress conditions in the rails are not met. For
a continuous deck, which is the most unfavourable condition, it is possible to reach
a jointless length in the range of 90 m. For very long structures more than one track
joint may be necessary.
6.6 Rail Expansion Joints 197

6.6 Rail Expansion Joints


6.6.1 Design of REJs – Calculation of the Maximum Displacement
Whenever an expansion device has to be installed, it shall be dimensioned depending
on its maximum displacement, thus selecting the appropriate type of device.
Usually allowed maximum expansion lengths are 300, 600, and 1200 mm. Even
though this maximum displacement capacity of 1200 mm is common, which would
represent a limit for maximum expansion lengths, new rail expansion joint (REJ),
with a maximum displacement capacity up to 1800 mm, are also starting to be devel-
oped and used.
For the dimensioning of the REJs, the displacements of the rail due to a variation
of temperature of deck and rail and due to shrinkage and creep shall be taken into
account. Usually the displacements due to braking/traction or bending can be dis-
regarded, except in special situations (very flexible fixed point or singular bridges).
The functional principle of the design of the REJs is based on the use of fixed
switch rails and moveable stock rails, as shown in Figure 6.10.
This kind of REJs is designed to assure a constant track gauge at all temperature
movements in the expansion. The rail track of the REJ comprises two half sets of
switches. One half set of switches consists of two stock rails and two flexible switch
rails, fully made from an asymmetrical tongue rail profile (see Figures 6.11 and 6.12).
Usually, the stock rails and the switch rails of the rail expansion joint will be
sup-ported by base plate systems provided with high elastic intermediate pads. The
de-sign of those intermediate pads will assure that vertical deflection of the base
plates within the rail expansion joint can be kept nearly constant for all base plates
regard-less of the shape, the dimensions of the plate and the number of rails to be
supported, which enable lower life-cycle costs for the rail infrastructure operator
in comparison with rigid or less elastic base plate systems, in which uncontrolled
vertical stiffness variations generate increasing dynamic loads and lead to rapid
degradations. If high elastic intermediate pads are used, the stock rails and the

Stationary part Mobile part of the bridge

Stock rail +
L

Switch
rail

Bridge
Stock rail
joint

Figure 6.10 Schematic depiction of the functional principle of REJs.


198 6 Longitudinal Track–Structure Interaction

72,03
5,3

14,56 Y

14.82
R200
R70

R200
R1

54
R90
6 1

3 R
R21 R2
R11

R1
1:20,17 R35 R35

1:20,17
R15

53,07
R3
R3
1:2, 75
75 1:2,
R19 R1
9
134,07

X X
1
28
R1
9
19
R

58,65
1:4

56
1:17
R4 44
+Y R4
20

20
3 R3
R +X
Y
55 2,67

140

Figure 6.11 Asymmetric switch rail profile 60E2A2 (Source: EN 13674-2).

Figure 6.12 Exemplary picture of a REJ (Source: Ferropedia).


6.6 Rail Expansion Joints 199

switch rails of the rail expansion joint are fixed by means of specially designed base
plates and rail-fastening components.
When the bridge has long expansion section lengths and thus large longitudinal
movements, it will be necessary at bridge joints to either reinforce the rail or provide
additional supports to the rail. It would not be necessary for REJs with a maximum
expansion length of 300 mm, but it will be for 600, 1200 mm, or longer.
The most common solution is the second option, i.e. maintaining the standard
rail section but providing additional supports to the rail in the bridge joint. Thus,
the decking inside the bridge gap is supported by steel moveable sleepers arranged
in accordance with the gap width (one for a maximum expansion length of 600 mm,
two for 1200, and three for 1800). These sleepers are mounted on auxiliary beams
aside from the track which act as auxiliary bridges. A crossbar control lever mech-
anism (see Figure 6.13) is used to adjust the position of these sleepers according to
the bridge gap movement so that the running rails are deflected comparable to the
other parts of the track.
With this mechanism the gap-opening variation is distributed over the moveable
sleeper spacings (between the last two sleepers at both sides of the joint and the steel
moveable sleepers) to ensure that:
The maximum spacing between sleepers is compatible with rail, fastening, and
the railway line characteristics: see Figure 6.14 with an example for a 1200 mm REJ,
in which maximum spacing remains below or equal to 650 mm, in the case of max-
imum opening of the lever mechanism of 1950 mm.
The minimum opening of the joint is limited: in the example of Figure 6.15 it is
higher than or equal to 190 mm in the case of minimum opening of the lever mech-
anism of 750 mm.
The maximum difference between one gap width between moveable sleepers and
the average gap is limited: in the example of this mechanism assures that the gap
with will vary between 650 mm (for the maximum opening of the mechanism of
1950 mm) and 250 mm (for the minimum opening of 750 mm).
On the other hand, the design of the bridge must assure that limit values for
maximum angular rotation and maximum vertical relative displacement between
adjacent structural elements (deck and abutment or adjacent decks) at joints are

Steel
moveable Longitudinal
sleepers girder

Ilanta

Structural Lever
joint mechanism
(scissors)

Figure 6.13 Lever mechanism for moveable steel sleepers (Source: [12]).
200 6 Longitudinal Track–Structure Interaction

650 mm 650 mm 650 mm 250 mm 250 mm 250 mm

1390 mm 190 mm

Maximum opening (winter) Minimum opening (summer)

Figure 6.14 View of the REJ’s gaps.

(a) (b)

Figure 6.15 Example of results of computations of uplift forces on fasteners (a) and
bending moments on rails (b) due to angular rotations and relative vertical displacements
of the decks.

fulfilled, as explained in Section 6.5. These verifications will assure that traffic safety
and comfort cannot be endangered by the creation of unacceptable changes in
vertical track geometry. However, in the presence of REJs, and specially with direct
fastened track, additional verifications must be done. Thus, it will be necessary
to verify: (i) that the REJ fasteners are not subject to unacceptable uplift forces;
(ii) that stress on rails and longitudinal girders of the REJ is acceptable.
The German Requirements Catalogue for the construction of slab tracks states that
the maximum tensile force on fastenings on both sides of the joint shall not be greater
than the force that decompresses the rail-pad, which depends on the fastening and
can be determined with normalised test of clamping force.
UIC Leaflet 776-3 states that in the case of direct fastened track on both sides of the
joint, angular rotation and vertical relative displacement must be limited to assure
that rail stresses are limited to 80 MPa, taking into account actual values of vertical
tensile stiffness of fasteners.
Some computations have been made that state that for usual values of vertical
tensile stiffness and force that decompresses the rail-pad, the limits established for
angular rotation and vertical relative displacement in the Eurocodes and UCI leaflets
will assure the fulfilment of these to verifications [13], but specific computations
must be made in every specific case.
6.6 Rail Expansion Joints 201

6.6.2 Regulation
Although the REJs are usually delivered at their neutral position, they shall be regu-
lated, at the moment of its welding, and installed at a position that takes into account
the expected remaining bridge movement (due to creep and shrinkage, forecasted
temperature variation, and breaking and traction traffic loads) and rail movements
(due to foreseen variations of rail temperature with regard to its neutral temperature,
as REJs shall be welded to the standard track at neutral temperature of the rail), and
breaking and traction traffic loads.
Although the devices can be adjusted if necessary, after being installed on the
track, in practice it is a very laborious process; therefore, it is highly recommended
to carry out the regulation before.
To do this, the REJ is first placed on a sliding surface (the ideal is to deposit it on
the rails of a well-levelled track or on the embedded rails of the turnout pre-assembly
slab installed in fabrication workshop or the mounting base).
In order to do this regulation of the REJs it will be necessary to measure the gap and
ambient temperature. It is advisable to make several measurements at different tem-
peratures and to do it few days before the regulation of the device (see Figure 6.16).
Based on these measures, it is possible to foresee significantly what will be the
dimension of the joint on the day of assembly at the scheduled time, especially if
there are very recent measures.
The state and position of the rail expansion device shall be recorded at installation
compared to a fixed reference point (FRP), by adjusting a pointing device that will
indicate the longitudinal movements of the bridge from the moment of installation.

6.6.3 Installation
They are installed on the side of the moving end of the structure. The fixed part of
the REJ must be on the fixed part of the viaduct (abutment – embankment), and the
movable part of the REJ on the superstructure of the viaduct itself (deck).
Once the position in which the element is to be placed has been defined, the guard
element is first laid out, and then the ballast is spread in a uniform layer.
This guard element is the necessary system for interrupting the ballast bed
(Figure 6.17a) to prevent the ballast from being deconsolidated by the movements
of the structure (which can occur if the ballast is continuous over the structural
joint) and to prevent it from falling through the joint opening. At the outer edges of
the wall there are wing walls to prevent the ballast from spilling over to the sides
in this area. The top of the wall must reproduce the final cant that the track will
have. An elastic sheet is usually placed between the two facing walls to catch loose
stones and other objects that may fall from the track for any reason. This sheet can
be folded and unfolded following the movements of the deck (Figure 6.17b).
Once the ballast bed has been prepared, the expansion device shall be positioned in
accordance with the assembly drawings and then moved into its theoretical position.
The flatness of the bearing assembly shall be maintained at all times.
202 6 Longitudinal Track–Structure Interaction

d = Distance between axles of sleepers nearest to the gap


dminimum = 750 mm dneutral = 1.350 mm dmaximum = 1.950 mm

Se = 790 − a
a (Tambient)

a
aminimum = 190 mm aneutral = 790 mm amaximum = 1.390 mm

Neutral situation before regulation


FRP and end of switch rail
The FRP matches the end of
switch rail
Switch rail
Stock rail

Abutment Deck
a ≥ 790 mm Se ≤ 0
End of switch rail Extended gap: the REJs must be
FRP open by moving the switch rail
Switch rail towards the fixed part
Stock rail

Abutment Deck
a ≤ 790 mm Se ≥ 0
End of switch rail
Reduced gap: the REJs must be
FRP
closed by moving the switch rail
Switch rail Stock rail towards the moveable part

Figure 6.16 Regulation of REJs.

(a) (b)

Figure 6.17 (a) Ballast guard walls; (b) elastic sheet preventing ballast stones falling
(Source: [12]).
6.7 Longitudinal Schemes 203

Then, immediately before welding, the expansion device is adjusted, maintaining


the distance calculated for the adjustment between the zero-point gauge framed in
the stock rail and the point of the switch rail.
Once the device has been adjusted, it shall be welded to the adjacent tracks.

6.7 Longitudinal Schemes


Track–structure interaction is a key requisite for selecting a structural type, espe-
cially as regards longitudinal continuity of deck, location of fixed points to coun-
teract temperature effects and fixed points for resistance to stress, and limitation of
displacements due to braking and traction.
There are two main options:
– continuous deck with fixed points that define dilatable and deformable length of
deck, where expansion devices may be eventually installed.
– statically determinate spans without longitudinal continuity, with fixed points on
each span, where interaction issues may be addressed without needing expansion
devices.
A continuous deck with only one fixed point could imply a long length submit-
ted to braking/acceleration actions. Taking this in mind, several types of resistance
schemes can be defined. The most used in non-seismic regions (the special case of
seismic regions will be treated in Section 6.7.6.1) are the following:
(A) Continuous deck with a single fixed point located at one of the abutments.
(B) Continuous deck with the fixed point located on one of the central piers.
(C) Simply supported spans without longitudinal continuity, with a fixed point on
each span.
(D) Mixed solution, with the deck divided in several continuous stretches each of
ones including several spans and one fixed point.
(E) Fixed points at the two abutments and a neutral central portal frame.
In Sections 6.7.1–6.7.6 several schemes will be explained in more detail, including
application examples.

6.7.1 Continous Deck with a Single Fixed Point Located at One of the
Abutments
6.7.1.1 General
This first solution can be applied for small, medium-size or even large deck lengths.
However, depending on deck length, the installation of expansion devices at the
other abutment could be necessary.
Usually, depending on the specific characteristics of the bridge, for lengths lower
than 100 m it is possible to avoid expansion devices.
On the other hand, for long bridges which need the installation of REJs at the
mobile abutment (Figure 6.18), it must be taken into account that the maximum
displacement of the device will increase with the expansion length of the deck. A
204 6 Longitudinal Track–Structure Interaction

REJ

Fixed point

Figure 6.18 Continuous deck with a single fixed point located at one of the abutments.

maximum displacement capacity of 1200 mm is common for REJs, which would


represent a limit for maximum expansion lengths. New REJs, with maximum dis-
placement capacity up to 1800 mm, are also starting to be developed and used.

6.7.1.2 Examples
Example 1. Bridge needing REJ
The example corresponds to a ballasted double-track continuous high-speed railway
bridge with 7 spans, measuring a total length of 165 m (Figure 6.19).
The deck consists of a voided concrete slab with a constant height of 1.90 m and a
width of 14 m.
The connection between the girder and each pier is performed using pot bearings
sliding in longitudinal direction. One of the abutments of the viaduct (the one on the
right in Figure 6.20) is firmly connected to the deck and is the only fixed point of the
viaduct; on the other abutment the bearings are also free in longitudinal direction.
Figure 6.20 shows the stresses in rail computed for a seasonal variation of temper-
ature of plus 35 ∘ C in the deck. As can be seen, only under the action of the variation
of temperature, the maximum compression in rail here is greater than the allowed
limit of 72 MPa for bridges with ballasted track.
Thus, it will be necessary to install an expansion device in the track to limit this
rail stress (Figure 6.21).
Once decided to install the REJ, it will be necessary to make the analysis including
all the actions to be considered.
The combinations used in this case consider, in a first step, seasonal variation of
temperatures combined with equivalent shrinkage and creep long-term deforma-
tion, produced after the installation of the CWR [13], which is equivalent to an
additional decrease of temperature of 35 ∘ C.
In a second step stresses for the vertical traffic forces are computed, considering
trains on both tracks, and for horizontal braking and traction forces, considering
several situations of trains breaking or starting.

165 m
20 m 25 m 25 m 20 m
4.70
1.90

14.00

Figure 6.19 Longitudinal scheme and deck cross section of bridge for example 1.
6.7 Longitudinal Schemes 205

100
80
60
Stresses in the rail

40
20
0
–20
–40
–60
–80
–100

Figure 6.20 Stresses in rail computed for a seasonal variation of temperature of 35 ∘ C in


the deck for example 1.

Expansion
165 m
device
20 m 25 m 25 m 20 m

Figure 6.21 Longitudinal scheme of example 1 with REJ at the mobile abutment.

Envelop
Tension limit = 105.0 MPa
Compression limit = –122.3 MPa Rail stresses (MPa)
250
Tension limit
200
150
100
50
0
–50
–100
–150
–200 Compression limit

Figure 6.22 Envelope of rail stresses computed for bridge of example 1.

Then the envelope of rail stresses for all the actions are computed. As can be seen
in Figures 6.22 and 6.23, those stresses are lower than the limit values (adding to the
limit values the rail stress due to rail temperature variations).
Also, the verifications of limit values for displacements are fulfilled.
206 6 Longitudinal Track–Structure Interaction

Displacements on the expansion device:


Displacements due to decrease of temperature:
Rail (maximum anual decrease): –50°C
Deck (equivalent to a maximum anual decrease plus deformation due to long term shinkage and creep): –70°C
Displacement (m)
0.06
0.04
0.02
0.00
–0.02
–0.04
–0.06 143 mm
–0.08
–0.10
–0.12
–0.14
Displacements due to increase of temperature: Maximum displacement = 239 mm.
Rail (maximum anual increase): +50°C
Deck (maximum anual increase): +35°C REJ required = 300 mm
0.12
0.10 Displacement (m)
0.08
0.06
0.04
0.02 96 mm
0.00
–0.02
–0.04
–0.06
–0.08

Figure 6.23 Computation of maximum displacement in the REJ for example 1.

And finally, to design the required expansion device its maximum displacements
are computed. In this case a maximum displacement of 239 mm is computed, so an
expansion device with an allowable displacement of at least 300 mm is required.
Example 2. Bridge not needing REJ
The usual one for this type of viaducts, with a fixed point in one of the abutments,
implemented by means of anchoring deck to abutment.
The example corresponds to a ballasted bridge of the high-speed line
Madrid–Barcelona. It is a continuous viaduct with four spans and an overall
length of 110 m, with independent decks for each track. The bearings system is the
same that for Example 1, using pot bearings sliding in longitudinal direction, and
with a connection between the deck and one of the abutments (the one on the right
in Figure 6.24) which is the only fixed point of the viaduct; on the other abutment
the bearings are also free in longitudinal direction.
First a computation has been executed considering continuous rail at the two
ends of the bridge, with a combination of variation of temperature of deck and rail,

110
20 35 35 20

E-1 E-2

P1 P2 P3

Figure 6.24 Longitudinal scheme for example 2.


6.7 Longitudinal Schemes 207

Maximum tensile = 202.1 MPa


Maximum compressión = 202.7 MPa Rail stresses
250
200
150
100
50
0
–50
–100
–150
–200
–250

Figure 6.25 Envelope of rail stresses computed for bridge of example 2.

deformation due to shrinkage and creep and traffic loads for different positions of a
train. Figure 6.25 shows the obtained envelopes of stresses in the rail. As can be seen
on the free abutment the limit of compression is slightly exceeded (72 MPa plus the
stress due to an increase of rail temperature of 50 ∘ C).
The situation which produces this maximum of compressions is due to the com-
bination of the following actions: decrease of temperature of deck equal to −58 ∘ C,
which is the sum of the seasonal maximum decrease of −35 ∘ C and the equivalent
temperature for the deformation due to shrinkage and creep of −23 ∘ C, and the train
located with its head on abutment number two and braking.
Considering that the limit of stresses is only slightly overpassed, but that the sim-
plified combination of actions used is usually conservative in terms of rail stresses,
for the same situation a more accurate step-by-step computation has been carried
out. To perform this computation, the first thing to do is to define a law for the lon-
gitudinal behaviour of the fastening–ballast system as a function of the vertical load,
as shown in Figure 6.26.
The maximum horizontal force F N will increase linearly with the vertical load on
the point, starting from a minimum value F 0 for a null vertical load, up to the max-
imum value of 60 kN/m according to standards for a maximum vertical load (that
can be considered as 80 kN/m per meter of distributed load according to the model
of UIC 71). The expression of F N will depend on the level of ballast maintenance as
shown in the figure.
Once these laws have been defined for non-linear springs the step-by-step compu-
tation proper can begin, where the first step shall be computing the situation after
deformations imposed in rail and deck.
After the first step the non-linear springs will be in different situations, as shown
in Figure 6.27, depending on the relative displacement that has occurred between
rail and infrastructure.
208 6 Longitudinal Track–Structure Interaction

t
N = 80 kN/m
60 kN/m

Increasing N

FN

N=0
F0

2 mm u

Moderate maintenance of ballast: Good maintenance of ballast:


F0 = 12 kN/m F0 = 20 kN/m
F80 = 60 kN/m F80 = 60 kN/m
μ = 0.60 μ = 0.75

Figure 6.26 Load-dependent force per length vs longitudinal displacement law.

Increase rail temperature (50 °C). Increase deck temperature (30 °C)
–60

–80

–100

–120

–140
Rail stresses (MPa)
–160
ΔT rail = 50 °C

ΔT in deck = 30 °C

N=0
F0

2 mm u

Figure 6.27 First step of step-by-step computation (strains imposed on rail and deck).

From the situation after step 1, the vertical loads are introduced step by step to
represent the train running along the viaduct. For each one of these loading steps the
situation of non-linear springs will change depending on the vertical load and the
relative displacement between rail and infrastructure, according to the previously
6.7 Longitudinal Schemes 209

t
N2
F2 = FN2

N1
FN1
F1

F0 N=0
Δu1 Δ u2

2 mm u

Figure 6.28 Successive steps of step-by-step computation (vertical traffic load).

Braking

t
N2
FN2

Δu2
N1
FN1

Δu1

F0 N=0

2 mm u

Figure 6.29 Final step of step-by-step computation (braking load).

defined laws, with the corresponding increases of force and variations of stresses
in rail. For the last computation step the braking forces are entered at the desired
position. The new relative displacements will then imply a new state of non-linear
springs, with increase of force or not, depending on the initial situation (Figures 6.28
and 6.29).
210 6 Longitudinal Track–Structure Interaction

Figures 6.30–6.32 allow a comparison of results as regards stresses in rail due to


vertical loads, braking and finally the combination of all actions. As can be seen in
this case the step-by-step method gives compression values lower than the limiting
value, and the installation of a REJ could be avoided. In this case the normative
compliance on displacements has also been checked for both methods.

5 Bending Step-by-step Method


Stresses in rail (MPa) Simplified Method

–5

Vertical
load

Figure 6.30 Additional rail stresses due to vertical traffic load for example 2, computed
with simplified combination and with step-by-step methods.

20
Braking
Stresses in rail (MPa)
Step-by-step Method
10
Simplified Method

–10

Braking
–20

Figure 6.31 Additional rail stresses due to braking load for example 2, computed with
simplified combination and with step-by-step methods.
6.7 Longitudinal Schemes 211

–60
Combination
–80

–100
Stresses in rail (MPa) Step-by-Step Method
–120 Simplified Method
Compression Limit
–140

–160

–180

–200
Braking

ΔT in deck
= –58 °C
Vertical
load

Figure 6.32 Envelope of rail stresses for example 2, computed with simplified
combination and with step-by-step methods.

6.7.2 Continous Deck with the Fixed Point Located on One of the
Central Piers
6.7.2.1 General
In this case, deck shall be free from longitudinal restraint at abutments and at the
rest of the pier.
Again, depending on deck lengths, the installation of expansion devices at both
abutments can be needed. This solution also implies transmission of the whole lon-
gitudinal braking/traction action to one pier, without help from abutments.

6.7.3 Simply Supported Spans Without Longitudinal Continuity, with a


Fixed Point on Each Span
6.7.3.1 General
A common structural solution for railway bridges decks is the use of simply sup-
ported spans. This solution is very often associated to the use of prefabricated beams
and, even though it is possible to given them longitudinal continuity, at least for lon-
gitudinal action if not for bending, the most common solution is without continuity.
Usually, in low-seismicity areas, a 4-bearing system is installed, longitudinally
fixing the deck to pier with restricted bearings while the other end is free with slid-
ing bearings (see Figure 6.33). More efficient solutions can be designed for seismic
regions, as explained in Section 6.7.6.1.
The main advantage of this solution from the point of view of longitudinal
track–bridge interaction is that very long bridges can be designed without needing
REJs, as can be seen in the following examples.
212 6 Longitudinal Track–Structure Interaction

Fixed bearing

Sliding bearing

Free bearing

Figure 6.33 Usual bearing system for simply supported span in low-seismicity areas.

958 m
46 m 46 m 46 m 46 m

A–1 A–2

P1 P2 P18 P19

Figure 6.34 Longitudinal scheme and deck cross section of bridge for example.

6.7.3.2 Example
The example corresponds to a ballasted double track simply supported high-speed
railway bridge with 20 spans, and an overall length of 958 m (Figure 6.34).
The deck consists of a HP-40 prestressed concrete box with a constant height
of 3.80 m. Each span has a fixed neoprene support at one end and a sliding
neoprene–Teflon support at the other. The fixed support is the closest to abutment 1.
Thus there is a sliding bearing support at abutment 2.
The following figures show the stresses in rail computed for (from top to bottom):
(i) a seasonal increase of temperature; (ii) a seasonal decrease of temperature and
associated equivalent decrease due to creep and shrinkage; (iii) vertical traffic load
for several positions of the train; (iv) braking and acceleration forces for the same
positions of the train. Finally, the envelope of rail stresses is shown in the lowest
figure. As can be seen, limit values are not reached and, consequently, installation
of REJs is not needed.
Also, the verifications of limit values for displacements are fulfilled (Figure 6.35).

6.7.4 Fixed Points at the Two Abutments and a Structural Joint in the
Middle
6.7.4.1 General
This solution can be applied for long bridges when the following conditions exist:

– It is not possible the solution described in Section 6.7.1: with the fixed point at
one of the abutments and the structural joint at the other, the expansion length of
6.7 Longitudinal Schemes 213

Rail stresses (MPa)


40
30
20
10
0
–10
–20
–30
–40
60
50
40
30
20
10
0
–10
–20
–30
–40
–50
–60
50
40
30
20
10
0
–10
–20
–30
–40
40
30
20
10
0
–10
–20
–30
–40
100 Tension limit
80
60 Tension
Compression
40
20
0
–20
–40
–60
–80 Compression limit
958 m
46 m 46 m 46 m 46 m
A–1 A–2

P1 P2 P18 P19

Figure 6.35 Rail stresses computed for bridge of example.

the bridge would be too high and, then, the expected movements on the structural
joint would be too large to be accommodated by a REJ.
– It is not possible the solution described in Section 6.7.2: the characteristics of the
foundations or the geometry (important height) of the central piers make impos-
sible to install the fixed point in the middle of the bridge.
214 6 Longitudinal Track–Structure Interaction

The alternative solution in these situations is the one presented in this section,
with fixed points at the two abutments and two continuous stretches of the deck,
with a structural joint, and then a REJ, in the middle of the bridge.
This is a very common solution in France, in which a neutral central portal frame
is used, combined with a double REJ (common in the French railways but not used
for other railway administrators.

6.7.4.2 Example
The example corresponds to a ballasted double-track high-speed bridge with 25
spans of lengths between 31.0 and 48.0 m, and an overall length of 1147.0 m. The
deck has two continuous stretches: section 1 between spans 1 and 14, with a length
of 685.0 m; section 2 between spans 16 and 25, with a length of 462.0 m. There are
structural joints at both ends of span 15, which becomes this way the neutral central
span (see Figure 6.36).
The deck consists on a HP-40 prestressed concrete box with a constant height
of 4.00 m. At both abutments there is a connection between deck and abutment
by means of prestressing steel. All bearings are sliding in longitudinal direction
neoprene–Teflon, except in the case of pier 14, corresponding to one of the ends of
neutral span 15 that are fixed in longitudinal direction.
A first computation of rail stresses is made without considering the presence of
rail expansion devices, for a seasonal variation of temperature of plus 36 ∘ C in the

E–1 E–2

P–1 P–5 P–9 P–18 P–21 P–24

P–14 P–15
Neutral span

Figure 6.36 Longitudinal scheme of the bridge.

Rail stresses (MPa)


100
50
0
–50
–100
–150
–200
–250
–300
E–1 E–2

P–1 P–5 P–9 P–18 P–21 P–24

P–14 P–15

Figure 6.37 Stresses in rail computed for a seasonal variation of temperature of 36 ∘ C in


the deck without REJs.
6.7 Longitudinal Schemes 215

deck. Figure 6.37 shows the stresses in rail computed. As can be seen, only under the
action of the variation of temperature, the maximum compression in rail is around
±240 MPa at the neutral span, much greater than the allowed limit of 72 MPa for
bridges with ballasted track. It is then clear the need of installing at least one rail
expansion device to limit this rail stress.
A second computation is made considering the presence of a REJ at pier 14 (mobile
end of section 1 and fixed point of neutral span). In this case it will be necessary to
make the analysis including all the actions to be considered.
The combinations used in this case consider: seasonal variation of temperatures;
equivalent shrinkage and creep long-term deformation produced after the installa-
tion of the CWR; vertical traffic forces considering trains on both tracks; horizon-
tal braking and traction forces, considering several situations of trains breaking or
starting.
Then the envelope of rail stresses for all the actions is computed. As can be seen in
Figure 6.38, those stresses are lower than the limit values (adding to the limit values
the rail stress due to rail temperature variations).
However, in this case the computed maximum displacement of REJ in this case
would be higher than 1200 mm. At the time of the design of this bridge REJs with an
allowed expansion length higher than 1200 mm were not available and this solution
was not possible (nowadays REJs for expansion length up to 1800 mm are used in
Chinese lines).
Computations were then repeated but considering a double REJ at the neutral
span, with rail joins at two ends of this span. This is actually the more usual design
for the track in the case of bridges with a neutral central span. As can be seen in
Figure 6.39, in this case again computed stresses are lower than the limit values.
In this case the maximum expansion computed for every of the two rail joints
is more reduced, but even though in the case of the joint on pier 15 is lower than

250
Rail stresses (MPa)
Tension limit
200
Tension
150 Compression
100
50
0
–50
–100
–150
–200 Compression limit

REJ

E–1 E–2

P–1 P–5 P–9 P–18 P–21 P–24


REJ

P–14 P–15

Figure 6.38 Envelope of rail stresses with REJ at pier 14.


216 6 Longitudinal Track–Structure Interaction

Rail stresses (MPa)


250
Tension limit
200
Tension
150 Compression
100
50
0
–50
–100
–150
–200 Compression limit

Double REJ

E–1 E–2

P–1 P–5 P–9 P–18 P–21 P–24


Double REJ

P–14 P–15

Figure 6.39 Envelope of rail stresses with a double REJ at neutral central span.

600 mm, in the case of pier 14 it is higher, and then it would be necessary to have a
device that allows a maximum expansion of 1200 mm (as explained in Section 6.6,
usual allowed maximum expansion length are 300, 600, and 1200 mm). However, a
600 mm device could be installed under special supervision and taking into account
that it would be necessary to carry out adjustment operations on it.

6.7.5 Deck Divided into Several Continuous Stretches, Each One


Including Several Spans and One Fixed Point
6.7.5.1 General
There are several situations in which this kind of solutions must be needed.
In the case of a very long bridge, in some situations it is not possible to install
fixed point for every span. Thus, the solution explained in Section 6.7.3, without rail
expansion devices, would be not possible.
In this situation, depending on the length of the bridge, it could happen that the
solutions with continuous deck explained in Sections 6.7.1, 6.7.2, and 6.7.4 lead to
too long expansion lengths. In this sense it must be considered that proven REJ com-
monly used (for instance in Europe) has an expansion capacity no larger than 1200
mm. On the other hand, larger REJs, with an expansion capacity of up to 1800 mm,
have been used in China and are being introduced in some project in Europe. In any
case, if the expansion length of the bridge leads to movement in the expansion joint
higher than 1220 or 1800 mm, it would be necessary to divide the deck into several
continuous stretches needing several REJs.
Alternatively, a solution with REJs can be also designed for this kind of situations.
It is the case when the fixed points can be installed, not of all the spans, but for
continuous stretches with expansion lengths not too high (around 100 m). This is
the case of the example included in this section.
6.7 Longitudinal Schemes 217

6.7.5.2 Example
The example corresponds to a double-track viaduct of a flying junction of a
high-speed line over a conventional existing railway line. To deal with the geometric
conflict between the position of the piers of the new viaduct and the tracks of the
conventional existing line, two different decks for every single track were necessary,
with lengths 630 m (left deck) and 608 m (right deck). The left viaduct consists of
21 spans, all of them 30 m long, whereas the right viaduct has 6 spans of 30 m,
10 spans of 28.5 m, one span of 23 m, and 4 spans of 30 m.
In both cases the deck consists in a HP-60 precast concrete girder with a depth of
2.0 m, and an in situ concrete slab HA-35 with a 0.25 m depth.
All the spans are simply supported. However, due to the geometrical difficulties
mentioned before, some of the piers do have not enough stiffness to become a fixed
point. Thus, the solution explained in Section 6.7.3 is not possible and it was nec-
essary to provide with axial continuity to some spans, by prestressing the joints, to
constitute continuous stretches in terms of longitudinal behaviour. The continuous
sections are:
Left deck:

– Section 1: spans 7, 8 and 9, with fixed point at pier 8


– Section 2: spans 10, 11, and 12, with fixed point at pier 10.

Right deck:

– Section 1: spans 7, 8, and 9, with fixed point at pier 8.


– Section 2: spans 10, 11, 12, and 13, with fixed point at pier 11.
– Section 3: spans 14, 15, and 16, with fixed point at pier 14.

In the case of spans without longitudinal continuity, there is a fixed point at one
end of each span. In these cases, the piers consist of a square-shaped pier shaft
founded on a four-pile cap.
In the cases of continuous sections, most of the supports consist on a portal frame
with two cylindrical piers, each of them founded on a single pile, under an HP-40
prestressed concrete beam with an inverted T section to support the decks. These
supports have not the required stiffness to become fixed points. The support for fixed
points is then constituted by a portal frame with two inclined piers on each side of
the existing conventional railway line, under the same HP-40 prestressed concrete
beam (see Figure 6.40).
Rail stresses are computed for the two viaducts without considering the presence
of rail expansion devices, for the following actions: seasonal variation of deck tem-
perature combined with equivalent shrinkage and creep long-term deformation pro-
duced after the installation of the CWR; vertical traffic forces considering trains on
both tracks; horizontal braking and traction forces, considering several situations of
trains breaking or starting. Results of the envelope of rail stresses for all the actions
are shown in Figures 6.41 and 6.42 for both viaducts.
As can be seen, limit values are not reached and, consequently, installation of REJs
is not needed.
218 6 Longitudinal Track–Structure Interaction

Pot
Continuity bearings
prestressing

P9 P10 P11 P12 P13

Figure 6.40 Longitudinal scheme of continuous stretch for spans 11 to 13 of right deck,
with fixed point at pier 11.

120 Rail stresses (MPa)


Tension limit
80 Tension
Compression
40

–40

–80 Compression limit


E1 P1 P2 P3 P4 P5 P6 P7 P8 P9 P10 P11 P12 P13 P14 P15 P16 P17 P18 P19 P20 E2

–120

Figure 6.41 Envelope of rail stresses for left viaduct.

120 Rail stresses (MPa)


Tension limit
80 Tension
Compression
40

–40

–80 Compression limit


E1 P1 P2 P3 P4 P5 P6 P7 P8 P9 P10 P11 P12 P13 P14 P15 P16 P17 P18 P19 P20 E2
–120

Figure 6.42 Envelope of rail stresses for right viaduct.

6.7.6 Especial Situations


6.7.6.1 Seismic Design
In high-seismicity zones, almost all bridge components are designed considering
seismic loads. Furthermore, if compared with viaducts in non-seismic zones, design
6.7 Longitudinal Schemes 219

Stiffness of fixed points of deck


means higher structure vibration
natural frequencies.

Elastic Response Spectra


0.80
Spectral Acceleration (g)

High natural frequency


Heavier bridges due to
0.60 requirements of in-
service deformability
0.40

0.20
Greater seismic
0.00 forces
0 1 2 3 4
Period (s)

Figure 6.43 Influence of the requirements on stiffness for fixed points on seismic design.

considerations have an impact not only on the volume of material but may also rad-
ically influence structural design.
Track–structure interaction design considerations are stricter for seismic condi-
tions. To ensure proper viaduct structural behaviour against braking/traction hor-
izontal actions, limiting deck displacements and stresses on rail, fixed points must
be established with sufficient stiffness. The requirements on stiffness for fixed points
mean higher structure natural frequencies in longitudinal direction and thus greater
seismic factors than for railroad bridges. In the case of road bridges, the use of seismic
isolator, as elastomeric bearings, usually leads to natural frequencies on the lower
zone of the spectrum, whereas in the case of railway bridges natural frequencies will
usually be on the zone of higher ground accelerations. That, combined with heavier
decks due to requirements of in-service deformability, makes that seismic forces will
be much higher (Figure 6.43).
Therefore, seismic considerations may be most important, especially for long
viaducts and even restrict the type of viaduct.
The seismic design of a railway viaduct should be based on a criterion of
seismic stress balance over piers and abutments compatible with requirements
of track–structure interaction. Longitudinal action is the most difficult analysis,
especially for long viaducts, due to the coupling of three actions of different nature:
(i) the seismic action itself; (ii) strains imposed by temperature variations and
rheologic effects; (iii) braking/traction actions. Thus track–structure interaction
is a key requisite for selecting a structural type, especially as regards longitudinal
continuity of deck, location of fixed points to counteract temperature effects and
fixed points for resistance to stress and limitation of displacements due to braking
and traction. Main variables will thus be bridge dilatable length(s), and deformable
deck length, depending on longitudinal continuity of deck and fixed-point location,
where the difference between dilatable and deformable length is that fixed points
may not be the same for slow deformations (temperature, rheologic effects) or for
fast ones (braking/traction, seismic action) because of installed STUs or dampers.
220 6 Longitudinal Track–Structure Interaction

The three main options of design exposed in Section 6.7 must be considered:

– Continuous deck with fixed points that define dilatable and deformable length of
deck, where expansion devices may be eventually installed.
– Simply supported spans without longitudinal continuity, with fixed points on
each span, where interaction issues may be addressed without needing expansion
devices.
– Mixed solution, with the deck divided in several continuous stretches each of ones
including several spans and one fixed point.

However, for high-seismicity zones, deck continuity issues should consider seis-
mic actions and their coupling with braking/traction and imposed strains, and also
relative displacements of deck.

Continuous Deck A continuous deck implies a large mass fully excited by seismic
action (deck plus total viaduct dead weight). Longitudinally, there are four types of
resistance schemes to the seismic force associated with the vibration of this mass:

(a) The first solution is the one already exposed in Section 6.7.1, usually used for
low-seismicity areas or small deck lengths, with a single fixed point located at
one of the abutments. This is also usually the chosen with high piers that can’t
become fixed points. As already explained this solution implies, depending
on deck length, the installation of expansion devices at the other abutment.
The main drawback is that in high-seismicity areas the fixed abutment shall
support the whole seismic action.
This is usually the case when we have high piers that can’t become fixed points.
(b) A second solution would be similar to the one explained in Section 6.7.2, in
which fixed point is located on one of the central piers. In this case, deck shall
be free from longitudinal restraint at abutments, which would imply transmis-
sion of the whole longitudinal action (seismic plus braking/traction) to the pier
in which the fixed point is situated.
Due to the need of high stiffness of the fixed point, the central pier used to
establish it is usually an especial one, a very rigid A-type pier or similar, and
with a large foundation, as in the examples of Figure 6.44.

Figure 6.44 Examples of especial central pier for fixed point in Spanish HS railway.
6.7 Longitudinal Schemes 221

Additionally, it must be considered that usually expansion devices will be


needed at both abutments.
The main drawback of these solutions with one fixed point is that it shall sup-
port the whole seismic action, so it is not valid for high-seismicity areas or long
viaducts. In this case the action shall be spread over a maximum number of
piers for better distribution of forces; thus, two variants can be implemented.
(b.1) Direct transmission of longitudinal forces to all or a few piers, implying restric-
tion on deck movement along most of deck and thus large stresses appearing
due to imposed strains (temperature variations and shrinkage/creep)
(Figure 6.45).
In this case, it would be necessary to find an optimum ratio between distribu-
tion of seismic and long-term deformation, as standards allow not to consider
them as concomitant.
In this solution the ductility of the piers can be used, allowing the appearance
of plastic hinges at the base section of central piers with fixed points under
design seismic action. Plastic behaviour is designated at the bottom of pier for
easy inspection and repair. With this solution reduced seismic force can be con-
sidered in the design of piers, whilst other members are designed by unreduced
seismic forces (Figure 6.46).

Piers with fixed bearings:


search for an optimum

Figure 6.45 Continuous deck with several fixed points located at the central piers.

Zone designated for the appearance of


plastic hinges under design seismic action

Figure 6.46 Appearance of plastic hinges under design seismic actions.


222 6 Longitudinal Track–Structure Interaction

(b.2) Transmission of longitudinal forces to all or a few piers, installing STUs


(Shock Transmitter Units) on all or a few piers supporting longitudinal forces
(Figure 6.47).
STUs allow non-restrained long-term movements avoiding stresses due to
imposed strains. And then the distribution of seismic action may be made
on a larger number of piers. But it must be considered the high cost of STUs
(installation and maintenance costs) and the doubts about its efficiency
against breaking actions.
(c) The following option would be transmission of longitudinal forces to abut-
ments installing STUs only at these points, bearing in mind that these devices
don’t dissipate energy and just block the seismic action which will be fully
transmitted to abutments. So, seismic force balance over two abutments may
be too high for high-seismicity areas or long viaducts (Figure 6.48).
This solution must be combined with some elastic bearings (not all the rest
can be sliding bearings) to assure re-centring of the deck after longitudinal
movements. STU performance against braking force is such that part of this
action is transmitted to piers through elastic bearing too.
(d) A similar option, in the case of high-seismicity areas, is installing dampers
for absorbing part of the seismic action, with consequent reduction of seismic
forces.
This solution can result in different variants, dampers only on abutments, on
abutments and piers, always associated with an adequate bearing system (to
deal with braking/starting forces).

Shock Transmitter Units

Figure 6.47 Continuous deck with several fixed points located at the central piers and
STUs on all or a few piers.

Elastic bearings
Shock Transmitter Units
for re-centering

Figure 6.48 Continuous deck with STUs located on both abutments.


6.7 Longitudinal Schemes 223

Dampers Elastic bearings


for re-centering

Figure 6.49 Continuous deck with dampers located on both abutments.

Figure 6.50 Arroyo de las Piedras Viaduct, Córdoba-Málaga HSL, Spain.

It is the case of the viaduct showed in Figure 6.49, in which dampers were
installed on both abutments.
(e) Finally, the deck can be segmented into segments longer that one single span,
which is an intermediate situation between continuous deck and simply sup-
ported. This solution can result in different variants, dampers only on abut-
ments, on abutments and piers, always associated with an adequate bearing
system (to deal with braking/starting forces).
It is the case of the viaduct showed in Figure 6.50, in which dampers were installed
on both abutments.

Discontinuous Deck The most common situation for discontinuous decks is the case
of simply supported spans without longitudinal continuity in which fixed points
with a sufficient stiffness will be necessary on each span.
Usually, in low-seismicity areas, a four-bearing system is installed, longitudinally
fixing the deck to pier with restricted bearings while the other end is free with sliding
bearings, as shown in Figure 6.33.
However, for high-seismicity areas this bearing system has some disadvantages.
In the first place, the presence of different bearings restricting the same deck
displacement, as here, with a fixed bearing and a sliding bearing in transversal
direction, because they have different tolerances, may produce a longitudinal load
imbalance that can be critical in the design of the bearings, especially against
seismic forces [14, 15] (Figure 6.51).
224 6 Longitudinal Track–Structure Interaction

k (stiffness)

d
t (tolerance)

Figure 6.51 Pot-bearing behaviour model.

Forces on bearings Relative displacement of piers

Deck rotation Deck rotation

Figure 6.52 Appearance of longitudinal forces on bearings due to relative lateral


displacements of piers under seismic action.

Fixed bearing

Rotation of deck Guided bearing


allowed
Free sliding bearing

Figure 6.53 Bearing system for simply supported span in high-seismicity areas.

On the other hand, transverse relative displacements between successive piers


may also produce deck rotations about the vertical axis. Four-bearing conventional
systems imply clamping between deck and pier; therefore, these rotations will in
turn induce vertical-axis moments on piers and longitudinal forces on bearings, with
probable impact on viaduct design and cost [14, 15] (Figure 6.52).
An improvement on bearing system (shown in Figure 6.53) may allow vertical axis
deck rotations, thus avoiding added stress due to transverse seismic action.
Finally, in all the previous schemes relative displacement between contigu-
ous decks is only restricted by the stiffness of pier and continuous rail. Due to
track–structure interaction limitations on this relative displacement, pier head
stiffness must be sufficiently large.
6.7 Longitudinal Schemes 225

Furthermore, in high-seismicity areas, although not required by standards, it is


usual considering two levels of seismic event:

i. Type I (severe or ultimate earthquake): design for repairable damages: struc-


tures are allowed to respond into the inelastic range with a ductility ratio not
exceeding the allowable demand so that all damage is repairable. Design is gen-
erally based on capacity design.
ii. Type II (moderate or service earthquake): design for safe operation at maxi-
mum speed and no yielding: to ensure that trains on the viaduct can brake safely
to a stop from full design speed: yielding of structural elements is not permitted
and track displacements must remain within the track–bridge interaction limit-
ing values with the structure subjected to seismic forces: no major damage to the
track is allowed.

This is the case of several HS lines, as can be seen in Table 6.3, as for instance
the Taiwan HSL, in which a Type I earthquake was considered for a return period
of 950 years and PGA up to 0.4 g, and a Type II service earthquake with a third
of the previous PGA, or the California HSL, in which a Maximum considered
earthquake is computed for a return period of 950 years, and an Operating Basis
Earthquake for a return period of only 50 years. And similar values for France and
Morocco.
Track–bridge interaction analysis under Service seismic action must be performed
to confirm that relative displacement is under the limits. However, as explained
in Section 6.3.1.3, the non-linear nature of the rail–structure interaction makes
time-history analysis unavoidable, as modal-spectral analysis is not possible. Addi-
tionally, longitudinal law is also dependent on the track being loaded or unloaded,
which further complicates the computation when combining seismic and traffic
actions (Figure 6.54).
In any case, synthetic displacement or acceleration time histories must be gener-
ated and applied at the base of the pier foundations for interaction analysis.
This problem can be solved by using STU, to provide the deck with structural conti-
nuity against seismic actions, to avoid relative displacements between adjacent decks
during earthquakes (Figure 6.55).

Table 6.3 Earthquake intensity levels.

USA
Country Taiwan (California) France Morocco

Ultimate earthquake RP = 950 years RP = 950 years RP = 475 years 1000 years
(type I) PGA = 0,4 g
Reduced earthquake RP = Undefined RP = 50 years RP = 20 to 50 Undefined
(type II) PGA = 1/3 years (0.582 ultimate
ultimate acceleration)
acceleration
226 6 Longitudinal Track–Structure Interaction

t
N = 80 kN/m
60 kN/m

increasing N
FN = F0 + μN
FN

N=0
F0

2 mm u

Figure 6.54 Load-dependent force per length vs longitudinal displacement law.

Fixed bearing

Rotation of deck Guided bearing STU


allowed
Free sliding bearing

Figure 6.55 Improved bearing system for simply supported span in high-seismicity areas
with STUs.

If these bearings also transmit horizontal actions (seismic plus braking/traction)


to piers at both ends of deck, relative displacements between contiguous decks are
controlled. If one of these bearings allows slow displacements (due to temperature
variations or shrinkage/creep), dilatable length remains unchanged and continuous
rail can be kept without expansion devices without excess stress on rail.

6.7.6.2 Exceptional Geometries


Usually for the analysis of conventional viaducts 2D models as the one shown in
Figure 6.4 are used. This type of models allows computing stresses and displace-
ments on conventional viaducts with mainly a 2D longitudinal behaviour, but it
is not suitable for exceptional geometries in which the longitudinal and transverse
deck deformations influence both rail stress and relative displacements between rail
and deck.
On the other hand, for this kind of especial geometries, there can be situations
where one track is on the deck and the other is still on the embankment and thus
they will behave differently.
For these cases the models must include geometry on plan of the structure and the
actual position of tracks.

6.7.6.3 Example of Exceptional Geometry


The example corresponds to the high-speed pergola bridge whose schematic layout
is shown in the following figure.
The pergola is composed of 94 prefabricated double T prefabricated concrete
beams, with a span between supports of 28.3 m, and in situ concrete compression
6.7 Longitudinal Schemes 227

Figure 6.56 Schematic layout of the viaduct.

Linear springs
Section type 1

Section type 2
Section type 1

Section type 1
Section type 2
Section type 1
in both directions Axis of track
for neoprenes Section of rails

Section type 5
Section type 6
Section type 5
Section
type 3

Section
type 4

Section
type 3

Non-linear springs in the rail direction


Rail-platform connection or rail-deck

Figure 6.57 Scheme of the model used for track–bridge interaction.

slab of variable thickness between 0.25 and 0.39 m. All the beams have laminated
elastomeric bearings of variable dimensions (Figure 6.56).
This is a clear case in which the use of a conventional 2D models is not suitable,
and a model as the one shown in Figure 6.57 is needed, to consider both longitudinal
and transverse deck deformations.
A simplification is made assuming that the two rails of each track can be consid-
ered as concentrated on the track axis has been made, as shown in the figure. In
this scheme it is observed that only the pergola part covered by an upper slab with
continuity throughout the width of the pergola has been idealised (see Figure 6.57),
which is a sufficient approximation to the deformation behaviour of the deck.
On the other hand, the interaction analysis takes into account the vertical loads in
order to assess the influence of deck bending on rail stresses. However, in this case,
the main direction of bending is the same as the beams and is almost at right angle
with track axis and therefore this effect can be neglected. Thus, the computation
model can be greatly simplified as a plane model can be sufficient. On one hand
strains and displacements can be considered on the structure plane and furthermore
all elements can be considered on the same plane without considering the different
228 6 Longitudinal Track–Structure Interaction

Figure 6.58 Case example: scheme of position of braking and acceleration forces on the
tracks.

Rail stresses (MPa)


100 Tension limit
80 Left track +
60 Left track –
40 Right track +
Right track –
20
0
–20
–40
–60
–80 Compression limit
–100

Figure 6.59 Envelope of rail stresses.

heights of bearings, axis of deck, deck upper deck and track, as is usually done to
take into account effects of bending.
Another special consideration is related to longitudinal traffic load. The different
considered situation must differentiate the position of the train in the two tracks, as
shown in (Figure 6.58) for one of the considered cases.
Finally, as in the previous examples, seasonal variation of deck temperature in
both directions, longitudinal and transversal, is combined with equivalent shrinkage
and creep long-term deformation produced after the installation of the CWR. As
creep deformation is different in each direction, that leads to different final values
considered of deck temperature decrease. Thus, temperatures variation considered
are:

– Increase: DT = 35 ∘ C in both directions;


– Decrease (considering equivalent effect of creep and shrinkage): longitudinal
direction DT = −67 ∘ C; transversal direction DT = −58 ∘ C.

Results of the envelope of rail stresses for the two tracks considering all the actions
are shown in Figure 6.59.
As can be seen, limit values are not reached and, consequently, installation of REJs
is not needed.
6.8 Example of Track–Structure Interaction 229

6.8 Example of Track–Structure Interaction

The example of the bridge introduced in Chapter 3 is used here to illustrate


track–structure interaction. Since the length of the bridge is significant (368 m), it
is most likely that a track joint will have to be placed at abutment 2. This will be
verified in the development of the example. A load classification factor 𝛼 = 1.21 is
considered. The longitudinal gradient of the deck at abutment 2 is 1.5%.
The rail is modelled as a beam element with an area of 7.67 × 10−3 m2 and an
inertia of 3.06 × 10−5 m4 . A breathing length of 250 m is modelled on the access to
abutment 1. The following loads are considered:

– Uniform temperature in rail ±50 ∘ C.


– Uniform temperature in deck −18 ∘ C and +30 ∘ C (assuming T max = 45 ∘ C,
T min = −10 ∘ C, and T 0 = 15 ∘ C).
– Positive temperature gradient: 6 ∘ C.
– Negative temperature gradient: −5 ∘ C.
– Braking force equal to 20 𝛼 kN/m but limited to 6000 𝛼 kN.
– Traction force equal to 33 𝛼 kN/m but limited to 1000 𝛼 kN.
– UIC 71 train. The locomotive is not accounted for, to simplify the calculation. The
train is, therefore, modelled as a load equal to 𝛼 × 80 kN/m. The effects of this
simplification on the results are negligible.
– A strain of 450 × 10−3 due to creep and shrinkage, allowing for the strain that
occurs before the placing of the track.

The dynamic load factor can be neglected, as specified in EN 1991-2 6.5.4.3.

6.8.1 Verification of Stresses in the Rails


Figure 6.60 shows the variation in stresses in the rail due to temperature increase
in the deck for the case when there is no track joint. The increase in compressive
stresses is around 130 MPa, far over the 72 MPa allowed value. It is therefore clear
that a joint is necessary. In further calculations, therefore, a track joint will be
assumed to be placed at abutment 2.
Figures 6.61 and 6.62 show the effects of imposed deformations Figure 6.61 repre-
sents an increase in temperature combined with a positive temperature gradient and
rheological strains in the deck. This combination leads to an increase in the value of
the maximum compressive forces in the rail, even though there is a joint. Figure 6.62
represents the effect of a decrease in temperature combined with the effects of nega-
tive temperature gradient in the deck. In this case, tensile stresses over the deck are
smaller than the tensile stresses that develop in the rail over the ground.
Figures 6.63, 6.64, and 6.65 represent the effect of railway traffic on the rail stresses,
considering loading of odd spans, even spans, and the full structure. The maximum
stresses are obtained for the first condition since this maximises rotations at abut-
ment 1. The effect of traffic is relatively moderate as it contributes a maximum of
25 MPa to the rail stresses.
230 6 Longitudinal Track–Structure Interaction

100
Variation of stress in rail (MPa)

50
Free end
(no joint)
Fixed end
Bridge
0
–250 –150 –50 50 150 250 350 450 550

–50

–100

–150
x (m)
Stresses in rail Limit in compression Limit in tension Bridge position

Figure 6.60 Increase of stresses in rail due to temperature increase in the deck only.

225

175

125 Free end


Stress in rail (MPa)

(joint)
75
Fixed end
25 Bridge
–250 –150 –50 –25 50 150 250 350

–75

–125

–175

–225
x (m)
Stresses in rail Limit in compression Limit in tension Bridge position

Figure 6.61 Total stresses in rail due to increase in temperature in deck and rail plus
positive temperature gradient plus creep and shrinkage.

Figures 6.66, 6.67, and 6.68 represent the effect of the braking and traction
forces. They differ in the position of the trains, which are placed from abutment 1
towards abutment 2 in the first figure, centred on the bridge in the second, and
from abutment 2 towards the bridge in the last one. Differences are small but the
maximum stresses occur at abutment 1 when horizontal forces are acting from
abutment 1 towards the bridge. The maximum stresses are about ±11 MPa.
Finally, Figure 6.69 shows the envelope of rail stresses. Also represented separately
are the envelopes due to imposed strains (uniform temperature, temperature gradi-
ent, creep, and shrinkage), the vertical train load, and the braking and traction forces.
6.8 Example of Track–Structure Interaction 231

225

175

125 Free end


Stress in rail (MPa)

(joint)
75
Fixed end
25 Bridge
–250 –150 –50 –25 50 150 250 350

–75

–125

–175

–225
x (m)
Stresses in rail Limit in compression Limit in tension Bridge position

Figure 6.62 Total stresses in rail due to decrease in temperature in deck and rail plus
negative temperature gradient.

50

40

30 Free end
(joint)
20
Stress in rail (MPa)

Fixed end 10
Bridge
0
–250 –150 –50 50 150 250 350
–10

–20

–30

–40

–50
x (m)
Stresses in rail Bridge position

Figure 6.63 Total stresses in rail due to train load placed on odd spans.

In this case it is clear that the effect of imposed strains is the dominating factor. Also,
the maximum stresses comply easily with the set limits for total rail stresses, indi-
cating that a significantly longer bridge can be designed using a single track joint,
placed at abutment 2.

6.8.2 Verification of Horizontal Displacement at Abutment 2 Due


to Braking and Traction Forces
The displacement at abutment 2 due to breaking and traction forces is 5.14 mm. This
puts the relative displacement between deck and abutment slightly over the 5.00 mm
limit (see Section 6.5.2). A longer viaduct would, therefore, require shifting the point
232 6 Longitudinal Track–Structure Interaction

50

40

30 Free end
(no joint)
20
Stress in rail (MPa)

Fixed end 10
Bridge
0
–250 –150 –50 50 150 250 350
–10

–20

–30

–40

–50
x (m)
Stresses in rail Bridge position

Figure 6.64 Total stresses in rail due to train load placed on even spans.

50

40

30 Free end
(joint)
20
Stress in rail (MPa)

Fixed end 10
Bridge
0
–250 –150 –50 50 150 250 350
–10

–20

–30

–40

–50
x (m)
Stresses in rail Bridge position

Figure 6.65 Total stresses in rail due to train load placed on all spans.

of anchorage from abutment 1 to an intermediate longitudinally stiff pier (such as


an A-frame).

6.8.3 Verification of Horizontal Displacement at Abutment 2 Due


to Vertical Train Loads
The displacement at the top of the deck is determined from the longitudinal dis-
placement and the end rotations of the deck. Table 6.4 details the calculation of the
displacement of the top of the deck at abutment 2. The distance of the centroid of
6.8 Example of Track–Structure Interaction 233

20

Free end
10 (joint)
Stress in rail (MPa)

Fixed end Bridge


0
–250 –150 –50 50 150 250 350

–10
Stresses in rail - Forces towards Ab. 2
Stresses in rail - Forces towards Ab. 1
Bridge position
–20
x (m)

Figure 6.66 Total stresses in rail due to braking and traction forces acting from
Abutment 1 towards Abutment 2.

20

Free end
10 (joint)
Stress in rail (MPa)

Fixed end Bridge


0
–250 –150 –50 50 150 250 350

–10 Stresses in rail - Forces towards Ab. 2


Stresses in rail - Forces towards Ab. 1
Bridge position

–20
x (m)

Figure 6.67 Total stresses in rail due to braking and traction forces acting centred on the
bridge.

Table 6.4 Displacements of the top of the deck due to train live load (LL) at abutment 2.

Load combination Displacement (mm) Rotation (mrad) 𝜹H (mm)

LL on odd spans 2.24 (→) 0.267 ( ) 2.73


LL on even spans −0.775 (←) −0.852 ( ) −2.33
LL on all spans 1.47 (→) −0.587 ( ) 0.38
234 6 Longitudinal Track–Structure Interaction

20

Free end
10 (joint)
Stress in rail (MPa)

Fixed end Bridge


0
–250 –150 –50 50 150 250 350

–10 Stresses in rail - Forces towards Ab. 2


Stresses in rail - Forces towards Ab. 1
Bridge position

–20
x (m)

Figure 6.68 Total stresses in rail due to braking and traction forces acting Abutment 1
towards Abutment 2.

225

175

125 Free end


(joint)
Stress in rail (MPa)

75
Fixed end
25 Bridge
–250 –150 –50 –25 50 150 250 350

–75

–125

–175

–225
x (m)

Envelope of all actions Envelope of imposed strains Envelope of train load


Envelope of braking and traction forces Limit in tension Limit in compression

Figure 6.69 Envelope of stresses in rail.

the deck section from the top is 1.85 m. The maximum value of the horizontal dis-
placement is 2.73 mm, which complies with maximum limit of 8 mm for a model
accounting for the stiffness of the track (see Section 6.5.2).

6.8.4 Verification of Vertical Displacement at Abutment 2 Due


to Vertical Train Loads and Temperature Variations
Longitudinal displacements due to uniform temperature, temperature gradient, and
live loads can result in a relative vertical displacement between the deck end and the
abutment if there is a longitudinal gradient in the deck, since displacement of the
deck relative to be abutment will be horizontal (see Section 6.5.2). The calculations
References 235

Table 6.5 Calculation of vertical displacement between deck edge and abutment 2.

Displacement Rotation 𝜹H 𝜹V
Load comb (mm) (mrad) (mm) (mm)

Temperature increase + 110.1 (→) 0.176 ( ) 110.43 1.70


negative gradient
LL on odd spans 2.24 (→) 0.267 ( ) 2.73
Temperature decrease + −66.66 (←) −0.200 ( ) −67.03 −1.04
positive gradient
LL on even spans −0.775 (←) −0.852 ( ) −2.33

of this vertical relative displacement are determined in Table 6.5, accounting for the
longitudinal gradient at abutment 2 of 1.5%. The maximum value of the relative ver-
tical displacement is 1.70 mm, which is smaller than the limit of 2.00 mm for a design
speed larger than 160 km/h.

References

1 Calçaza, R., Delgado, R., Campos e Matos, A. et al. (2008). Track-bridge Interac-
tion on High-Speed Railways. Boca Raton: Taylor & Francis.
2 Granell, I., Arrieta, J.M., Celemin, J., et al. (2009). Ejemplos de Aplicación de la
IAPF-07. ACHE Monografía M-15.
3 Union Internationale des Chemins de Fer (2001). UIC Leaflet 774-3R.
Track-bridge interaction. Recommendations for calculations.
4 European Committee for Standardization (CEN) (2003). Eurocode 1 EN 1991-2.
Actions on structures – Part 2: Traffic loads on Bridges.
5 Union Internationale des Chemins de Fer (1979). UIC Leaflet 776-1R. Loads to
be considered in railway bridge design.
6 European Committee for Standardization (CEN) (2003). Eurocode 1 EN 1991-1-5.
Actions on structures – Part 1-5: General actions – Thermal actions.
7 Cuadrado Sanguino, M. and González Requejo, P. (2004). Consideración de las
deformaciones por retracción y fluencia en el estudio de fenómeno de interac-
ción vía-tablero en el proyecto de puentes ferroviarios. Revista de Obras Públicas
3446: 45–51.
8 Granel, I., Arrieta, J., Celemin, J. et al. (2009). Interacción Vía-Estructura en
puentes ferroviarios. Algoritmos de cálculo paso a paso. Revista de Obras Públicas
3499 (156): 39–48.
9 Cuadrado, M., González, P., and Losa, D. (2010). New considerations on
track-structure interaction in railway bridges. In: IABMAS2010: The Fifth
International Conference on Bridge Maintenance, Safety and Management,
Philadelphia.
236 6 Longitudinal Track–Structure Interaction

10 Cuadrado, M. and González, P. (2007). La interacción vía-estructura en puentes


ferroviarios. In: Jornadas técnicas internacionales: La ingeniería para Alta Veloci-
dad. Veinte años de experiencia en España, Córdoba.
11 Cuadrado, M. and González, P. (2007). Numerical methods for the analysis of
longitudinal interaction between track and structure. In: Track-Bridge Interaction
on High-Speed Railways, Porto.
12 Mendoza, J. (2012). Montaje de aparatos de dilatación en Adif . Madrid: Fun-
dación Caminos de Hierro.
13 Goicolea, J., Cuadrado, M., Viñolas, J. et al. (2013). Estudio del comportamiento a
medio y largo plazo de las estructuras ferroviarias de balasto y placa. Monografía
111. Madrid: CEDEX; Centro de Publicaciones, Ministerio de Fomento.
14 González, P., Cuadrado, M., and Palacios, G. (2008). A probabilistic approach of
the behavior of restricted bearings of railway bridge decks. Assessment of the
load increase as a function of the stiffness and tolerances. In: World Congress on
Railway Research, Seoul.
15 González, P., Cuadrado, M., and Losa, D. (2006). Deck bearings for railway
bridges. Assessment of load increase due to tolerances on restricted bearings. In:
14th International Symposium EURNEX, Zelina.
239

Conceptual Design for Maintenance


José Romo

7.1 Introduction
This section outlines some considerations that must be taken into account in the
structural design of the bridge in order to improve its durability and reduce mainte-
nance work during its service life.
Designers of such bridges should be acquainted of the most updated versions of
the catalogue of damages used by inspectors, in order to elaborate designs oriented to
avoid the defaults and deteriorations identified by experience. Furthermore, design-
ers must keep in mind the need of providing access to important elements, such
as anchor heads, bearing devices, or expansion joints. In particular, design must
take into account sufficient overall robustness in order not to depend on elements
or details that cannot be sufficiently inspected or monitored, such as hangers and
prestressing tendons, that could lead to brittle failure.
From a maintenance point of view, the optimal structures are those that are inte-
gral, i.e. where there are no support devices or expansion joints. However, except
for very short structures, it is not possible to build completely monolithic bridges.
Nevertheless, the trend in some countries, e.g. Germany [1], is to build semi-integral
bridges that concentrate in a few piers or abutments the points where the expansion
joints and bearings are located. It is also very important not to forget that for the gen-
eral maintenance of the construction site it is best to have the minimum number of
rail joints and this is only possible by having some expansion joints in the structure
every 120 m or so.
Once the design has been conceived in general terms, the following aspects must
be taken into account:
– Accesses to the various elements of the deck.
– Possibility to inspect and change bearings, expansion joints, or other structural
components especially susceptible to corrosion attack.

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
240 7 Conceptual Design for Maintenance

– Drainage system.
– Systems to avoid eddy currents.
All these aspects are developed in Sections 7.1–7.5.

7.2 Accesses
One of the most important aspects when considering the maintenance strategy is
to provide access to the various structural elements of the bridge. In principle, areas
hidden from inspection should be avoided and, therefore, the interior areas of decks
or piers and abutments should be accessible. It is also highly recommendable to
provide sufficient lighting so that inspectors can carry out their work properly.

7.2.1 Decks
The decks must be accessible from the inside. To this end, the corresponding
accesses must be arranged in such a way that any point of the deck can be accessed
without difficulty. The entrance can usually be made either from the abutment
or from the pile heads. It should be borne in mind that, in addition to allowing
inspection personnel to pass through, it is necessary to take into account the
possible material and equipment to be moved inside the deck: stressing jacks,
support devices, etc. (Figure 7.1).

7.2.2 Piers
The piers must be accessible from the top, not only to inspect the bearings, but also to
replace them. This may condition the dimensions of the pier head, since in addition
to the space needed to place the temporary jacks, the space needed to move them and
to remove and place the new bearings must also be taken into account. In addition,

B A

B A
SECTION A-A SECTION B-B

Figure 7.1 Example of access on a deck.


7.2 Accesses 241

C C

SECTION A-A SECTION B-B

B
SECTION C-C

Figure 7.2 Example of access on a pier.

in the case of hollow piers, the inside of the shaft must be accessible by means of an
internal access staircase (Figure 7.2).

7.2.3 Abutments
As in the case of piers, the bearings placed on abutments must be replaceable, so
the necessary access must be provided, both for personnel and for the handling of
the supporting devices and the jacks and other devices necessary for the dismantling
and assembly of the bearings.
242 7 Conceptual Design for Maintenance

0.05 1.80
0.40 1.00
5.25 0.50 5.00 0.40 0.40
0.85

0.35
0.30
0.47

0.30
0.47 0.85 10 %

Inspection
gallery
1.70

0.30
0.30

1.83
2%
Metal door for

0.30
1.25
access a hollow
gallery 170 x 85 cm. 0.90 1.25
on left wing

5.02
7.30 1.50 6.30 1.00 1.50
1.80

1.80

8.80 8.80
Elevation Section

Figure 7.3 Example of access chamber on an abutment.

In the case of rather horizontal anchoring elements (prestressing, shock trans-


mitters, etc.) the design has to take into account both their ease of inspection and
maintenance as well as their potential replacement. It is common for there to
be a chamber behind the abutment wall from which such elements are accessed
(Figure 7.3).

7.3 Bearings

In general, high-speed rail bridges tend to have high-quality bearings. Typical bear-
ings are of the POT or spherical type, while the normal neoprene strapped bearings
are generally not allowed.
As indicated above, it is usual that the replacement of the supports is carried out
by entering from the deck at the top of the pier. For this reason, it will be necessary
to have an opening in the bottom of the deck of sufficient size to give access to main-
tenance personnel and to allow the passage of the devices necessary to carry out the
replacement of the bearings.
The reinforcement layout of the top part of piers and abutments shall take into
account the position and loading transmitted by the auxiliary jacks used during the
substitution of bearings (Figure 7.4).
7.4 Expansion Joints 243

D D

C C

SECTION A-A SECTION B-B

Jacks for Jacks for


bearing bearing
B substitution
substitution

A A

Pots
bearings Pots
bearings
Jacks for Jacks for
bearing B bearing
substitution substitution
SECTION C-C SECTION D-D

Figure 7.4 Example of layout for bearing replacement.

7.4 Expansion Joints


The expansion joints in the deck are areas through which ballast can enter and are
also a passageway for water that can damage the top of the piers or the abutments, as
well as their adjacent walls. To avoid the former, it is necessary to provide steel plates
to prevent the ballast from falling (as known, ballast can impair proper functioning
of bearing devices) Figure 7.5. These plates are generally fixed to the abutment and
can slide over another plate at the end of the deck [2]. In addition, there must be an
element that channels the water (see Section 7.5) that enters through the joint and
244 7 Conceptual Design for Maintenance

Plate on kerb wall

A 200

Plate e = 15
in channel area
Back plate

500
Plate e joint Bolt Ø20
of dilation in Plate e = 15
moving abutment in passage area

Drill Ø80 mm
with grid to the
entry cada 2.00 m "Concrete in-situ"

Precast
Betun Elastomer membrane edge beam
Non-affecting passable bolts reinforced with geotextile (4 mm)
to the safety of the circulation
of work vehicles

A
Sidewalk detail

15 mm S = 0.20 m
e 20 mm 0.20 m = S = 0.40 m
25 mm 0.40 m = S = 0.60 m
30 mm 0.60 m = S = 1.00 m
Movement
Smax+600
Plate e Plate e = 15 100

Bolts Plate e = 10
200

Ø20/0.20
Bolts
Ø20/0.50
100 200 100 S 100

Deck Fixed
abutment

Joint on fixed abutment


SECTION A-A

Figure 7.5 Example of steel plates joint in a fixed abutment.

leads it to the point of accumulation or discharge of water. In cases where a structural


expansion joint coincides with a rail joint in the same section, no cover plates shall
be fitted. However, in this case neoprene joints may be used (Figure 7.6).
There are also commercial devices that ensure the correct functioning of this area,
which is so delicate for the durability of bridges. Some examples of such joints are
shown in Figures 7.7 and 7.8.
7.4 Expansion Joints 245

Plate on kerb wall

B 200
Plate e = 15
in channel area
Back plate
500

Plate e joint Bolt Ø20


of dilation in Plate e = 15
moving abutment in passage area

Drill Ø80 mm
with grid to the
entry cada 2.00 m "Concrete in-situ"

Precast
Betun elastomer membrane edge beam
Non-affecting passable bolts reinforced with geotextile (4 mm)
to the safety of the circulation
of work vehicles

B
Sidewalk detail

Movement+100 S

Ø100 Ø9 0
e = 4–6 mm. e = 4–6 mm.

Welding
Railing joint detail

15 mm S = 0.20 m
e 20 mm 0.20 m = S = 0.40 m
25 mm 0.40 m = S = 0.60 m
30 mm 0.60 m = S = 1.00 m
100 Movement L
Plate e Plate e = 15 100

Bolts Plate e = 10
200

Ø20/0.20 Bolts
Ø20/0.50
200 100 S 100

Deck Fixed
abutment

Joint on movable abutment


SECTION B-B

Figure 7.6 Example of steel plates joint in a movable abutment.


246 7 Conceptual Design for Maintenance

0.05
0.32 0.45 0.43

Protective Protective
concrete plate Bitumen grout

0.06
0.25 0.20
Sand filling
T 120, galvanized

0.10
Mortar MGIII

Filter stones
0.42

0.45 0.45 0.32 0.46


0.05

Figure 7.7 Example of neoprene expansion joint (MAURER SPS GmbH).

Figure 7.8 Example of expansion joint (MAURER SPS GmbH).

7.5 Drainage

Adequate drainage is the cheapest and most effective way of preventing degra-
dation of structures. In this regard, it is strongly recommended to follow the
indications of documents like, among others, the CEB Durable concrete structures
CEB Design Guide [3], regarding the appropriate selection of shapes, presence of
gutters, etc.
7.5 Drainage 247

In general, water will be channelled through closed pipes and drainage will be
through the piers (if they are hollow and can be inspected) and/or through the abut-
ments. Figures 7.9 and 7.10 show examples of the shape of the drainage network of
a bridge.
In general, drain troughs like the one drafted in Figure 7.9 are highly recom-
mended, since they avoid water dripping or running down the front walls of
abutments and deck (Figure 7.10).

DUCT Ø300 DUCT Ø300


Cleaning hatches Cleaning hatches

DUCT Ø300

Funnel

DUCT Ø300

Funnel

Figure 7.9 Example of drainage of a deck.

Tooth plate
Anchors

Drain trough

Elevation

Figure 7.10 Drain trough expansion joint (Source: [4]).


248 7 Conceptual Design for Maintenance

7.6 Conclusions
Maintenance must be included in the initial design phases of a bridge. In fact,
in many cases, maintenance needs can shape both the structural solution and
the dimensions of the different elements that make up the bridge. In any case,
the maintenance design of a high-speed railway bridge cannot be independent
from the general track maintenance strategy. In many cases the prioritisation of
track conservation, e.g. by eliminating the rail joint, leads to a higher number of
structural joints which are contrary to the maintenance of the structure. On the
other hand, the construction of long viaducts with a single expansion joint requires
track joints, which are always a critical point for track maintenance. Perhaps the
intermediate alternatives, as proposed in the Deutsche Bahn Design Guide for
Railway Bridges [1], are the best solutions when analysing the conservation of the
railway infrastructure as a whole.
In the authors’ opinion, no important structural element or detail must be hidden,
inaccessible to inspection, monitoring, and potential replacement. In this regard,
the designed lifespan of the overall structure must be compatible with the likely
shorter life of such elements, as well as with secondary but important components
like drainage pipes, gutters, and drain troughs.

References

1 Leitfaden Gestalten von Eisenbahnbrücken. 1. Aufl., Berlin: DB Netze AG, Dezem-


ber 2008.
2 Norma ADIF Plataforma NAP 2-0-0.1. Puentes y viaductos ferroviarios. 2a Edición
2019.
3 Durable Concrete Structures - CEB Design Guide – 2nd Edition, 1989.
4 Office of Bridges and Structures [Ed.] (2014) IOWADOT, Bridge Maintenance
Manual. https://siims.iowadot.gov/IowaDOT_BridgeMaintenanceManual.pdf.
249

Appendix A

Basic Concepts of Dynamics


Alejandro Pérez-Caldentey

A.1 Dynamics of Single Degree-of-Freedom Systems

A.1.1 Dynamic Response to Moving Loads (Dynamic Load Factor)


This annex is meant to provide the reader with the basic notions upon which the
analysis of the dynamic effects produced by traffic on railroad bridges is based. The
explanation of the Dynamic Load Factor relies heavily on the basis explained in
the classic text Introduction to Structural Dynamics [1]. The reader can find more
detailed information in this reference.
The response of a structure to a single load that is applied dynamically differs
considerably from its behaviour when subjected to a load that is applied at a very
slow rate so that the application can be considered static. The response is a function
of the ratio between the duration of the load and the structure’s first natural
period. The Dynamic Load Factor is a simplified way to account for the dynamic
effect of the load. To better develop this concept, the dynamic equilibrium of a
single degree-of-freedom (SDOF) system without damping will be considered. The
dynamic equilibrium of the freely vibrating system, depicted in Figure A.1, is given
by Eq. (A.1).
M ÿ + Ky = 0 (A.1)
The solution to this equation is, of course, the harmonic movement equation given
in Eq. (A.2), as can be easily verified.
√ √
K K
y = A cos t + B sin t (A.2)
M M
⏟⏟⏟ ⏟⏟⏟
𝜔 𝜔

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
250 Appendix A Basic Concepts of Dynamics

Factors A and B depend on the initial position of the system, with respect to its
position of static equilibrium y0 , and its initial velocity ẏ 0 as shown in Eq. (A.3):
⎧y = y → A = x
⎪ 0 0
t=0→⎨ ẏ 0
⎪ẏ = ẏ 0 → ẏ = −𝜔A sin 𝜔t + 𝜔B cos 𝜔t → B = 𝜔


y = y0 cos 𝜔t + 0 sin 𝜔t (A.3)
𝜔
When the problem is one of forced vibration, the move-
ment will depend on the load-time law. If it is assumed
M
that a load is introduced suddenly and maintained con-
stant in time, the dynamic equilibrium equation can be
written as in Eq. (A.4).
K
M ÿ + Ky = F (A.4)

For such an equation, the solution would be the sum of


the harmonic solution of the homogenous equation plus
a particular solution of the full equation. In this case the
Figure A.1 SDOF particular solution is just y = F/K, so that the solution to
system.
the full equation can be written as:

F
y = A cos 𝜔t + B sin 𝜔t +
K
⎧ F
⎪y = y0 → A = y0 − K
t=0→⎨ ẏ 0
⎪ẏ = ẏ 0 → ẏ = −𝜔A sin 𝜔t + 𝜔B cos 𝜔t → B =
⎩ 𝜔
( ) ẏ 0
F F
y = y0 − cos 𝜔t + sin 𝜔t + (A.5)
K 𝜔 K
For a system that is initially at rest and placed at the static equilibrium position
(i.e. y0 = 0 and ẏ 0 = 0), Eq. (A.5) simplifies to Eq. (A.6).
F
y= (1 − cos 𝜔t) (A.6)
K
⏟⏟⏟
yst

If plotted (see Figure A.2), it becomes apparent that for this case, the maximum
dynamic deflection reaches 2 times the static deflection yst = F/K. The ratio between
the dynamic deflection and the static deflection is known as the Dynamic Load
Factor (DLF or 𝛷). The case considered above, where the force is applied instan-
taneously and maintained indefinitely, is the worst situation for a non-periodic
load.
A more general expression may be obtained for the DLF when the load history is
less extreme. When the SDOF system is subjected to a dynamic load, F(t), having a
generic load-time law, this load can be seen as a superposition of impulses, that is,
forces that are applied during a very short time (see Figure A.3).
A.1 Dynamics of Single Degree-of-Freedom Systems 251

2.50

2.00

1.50
y/yst

1.00

0.50

0.00
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time, t (s)

Figure A.2 Dynamic over static deflection resulting from a SDOF system initially at rest
when a load is applied suddenly and maintained through time. Resulting DLF = 2.

Figure A.3 A generic F(t)


force-time law can be F(t) = Fmax f(t)
decomposed into a series of Fmax
impulses.

t=τ

t

The solution of the dynamic equation for a single impulse is straightforward. If the
impulse is assumed to be applied when y = 0, and the duration of the load is very
small (as per definition of impulse), it can be assumed that during its application
the y coordinate does not change, so the equilibrium equation simplifies as shown
in Eq. (A.7):
F F
M ÿ + K y = F → ÿ 0 ∼ = constant → ẏ 0 = d𝜏 (A.7)
⏟⏟⏟ M M
∼0

The general solution given in Eq. (A.3) then simplifies to Eq. (A.8):

Fd𝜏
y= sin 𝜔t (A.8)
M𝜔
252 Appendix A Basic Concepts of Dynamics

If the impulse is applied at a time 𝜏, Eq. (A.8) will have a difference in time phase
as shown in Eq. (A.9):
Fd𝜏
y= sin 𝜔(t − 𝜏) (A.9)
M𝜔
Assuming linear elastic behaviour, the solution of a SDOF system subjected to
an arbitrary load, varying in time, can be obtained by superposition of a series of
impulses as shown in Eq. (A.10):
t t
F(𝜏) Fmax
y= sin 𝜔(t − 𝜏)d𝜏 = f (𝜏) sin 𝜔(t − 𝜏)d𝜏
∫0 M𝜔 M 𝜔 ∫0
⏟⏟⏟
K
𝜔2

t
Fmax
= 𝜔 f (𝜏) sin 𝜔(t − 𝜏)d𝜏
K ∫0
⏟⏟⏟
yst

t
y
→ = 𝜔 f (𝜏) sin 𝜔(t − 𝜏)d𝜏
yst ∫0
( )
y
𝛷 = max (A.10)
yst
This equation can be easily generalised to account for damping. The general
expression in this case is given in Eq. (A.11):
t
y
𝛷= = 𝜔d f (𝜏)e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
yst ∫0

𝜔d = 𝜔 1 − 𝜉 2 (A.11)

where:
𝜉 is the damping index (ratio between the structural damping and the critical damp-
ing1 ), which, for reinforced concrete structures, is around 2%.
With this expression the DLF(𝛷) can be determined for other situations. As an
example, the DLF(𝛷) for a constant load that is applied for a time equal to 1/4 times
the period of the structure can be solved as follows:
The SDOF system will be subjected to two regimes:

– The first one is valid while the constant load is applied. In this case, Eq. (A.6) is
fully applicable:
T
y = yst (1 − cos 𝜔t) t< (A.12)
4
Of course, in this case, the maximum is reached for t = T. The maximum value of
this expression for t < T/4 corresponds to t = T/4 with y = yst .

1 Value of damping which, if equalled or surpassed, results in a non-oscillating system.


A.1 Dynamics of Single Degree-of-Freedom Systems 253

– The second regime applies when t is greater than T/4, when the system vibrates
freely with the initial conditions provided by the first regime: initial position
(y0, 2nd ) and initial velocity (ẏ 0, 2nd ):
( ) ( )
T 2π T
y = yst 1 − cos = yst
(4) T 4
T 2π T
ẏ = yst 𝜔 sin = yst 𝜔
4 T 4
ẏ 0, 2nd
y = y0, 2nd cos 𝜔t + sin 𝜔t
} 𝜔
y0, 2nd = yst
→ y = yst (cos 𝜔t + sin 𝜔t) (A.13)
ẏ 0, 2nd = yst 𝜔
The maximum
√ of this expression is attained for 𝜔t = π/4, for which the deflection
would be 2yst , so the DLF would be 1.41.
It is easy to repeat this calculation assuming different durations for the application
of the force, td /T, (or different impulses). The result of this analysis is summarised
in Figure A.4. When the duration of the load reaches 0.5T, the DLF becomes equal
to its maximum value (i.e. 2.0). It can also be noted that when the duration of the
load is short, then the DLF can be smaller than 1.0.
Another case with a practical application for the design of railway bridges is the
triangular pulse. Figure A.5 shows an example of a triangular pulse of 380 kN ampli-
tude. The problem can again be solved by using Eq. (A.10). In this case, four time
intervals need to be distinguished:
– Before the impulse is applied, (t < t0 ) at which time the response of the system will
be nil
– The interval t0 ≤ t ≤ t0 + Δt/2, during which time the value of the force is
increasing

2.50

2.00

1.50 √2
DLF = y/yst

1.00

0.50

0.00 0.25
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
td /T

Figure A.4 DLF as a function of load duration (constant load).


254 Appendix A Basic Concepts of Dynamics

400

350

300

250
F(t) (kN)

200

150

100

50

0
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90
t (s)
Single triangular pulse

Figure A.5 Triangular pulse applied at time t 0 with a duration Δt.

– The interval t0 + Δt/2 ≤ t ≤ t0 + Δt, during which time the value of the force is
decreasing
– After t0 + Δt when the system will behave as a freely vibrating system with initial
conditions determined by the position and the velocity of the system for t = t0 + Δt.

For the second time interval,


t
F(𝜏)
y(t) = sin 𝜔(t − 𝜏)d𝜏
∫0 M𝜔
Δt
if t0 ≤ t ≤ t0 +
2
t
𝜔2 2(𝜏 − t0 )
y(t) = yst sin 𝜔(t − 𝜏)d𝜏
𝜔 ∫t0 Δt
( t t )
2𝜔
= yst 𝜏 sin 𝜔(t − 𝜏)d𝜏 − t0 sin 𝜔(t − 𝜏)d𝜏 (A.14)
Δt ∫t0 ∫t0
The first integral can be solved by integrating by parts:
b b
uv′ d𝜏 = uv − u′ vd𝜏
∫a ∫a
with:
1
u = 𝜏 and v′ = sin 𝜔(t − 𝜏) → u′ = 1 and v = cos 𝜔(t − 𝜏)∶
𝜔
b [ ]b b
𝜏 1
𝜏 sin 𝜔(t − 𝜏)d𝜏 = cos 𝜔(t − 𝜏) − cos 𝜔(t − 𝜏)d𝜏
∫a 𝜔 a 𝜔 ∫a
[ ]b
𝜏 1
= cos 𝜔(t − 𝜏) + 2 sin 𝜔(t − 𝜏) (A.15)
𝜔 𝜔 a
A.1 Dynamics of Single Degree-of-Freedom Systems 255

Therefore:
t
F(𝜏)
y(t) = sin 𝜔(t − 𝜏)d𝜏
∫t0 M𝜔
( t t )
2𝜔
= yst 𝜏 sin 𝜔(t − 𝜏)d𝜏 − t0 sin 𝜔(t − 𝜏)d𝜏
Δt ∫t0 ∫t0
([( ) ]t )
2𝜔 𝜏 − t0 1
= yst cos 𝜔(t − 𝜏) + 2 sin 𝜔(t − 𝜏)
Δt 𝜔 𝜔 t0
(( ) ( )
2𝜔 t − t0 t − t
= yst cos 𝜔(t − t) − 0 0 cos 𝜔(t − t0 )
Δt 𝜔 𝜔
)
1 1
+ 2 sin 𝜔(t − t) − 2 sin 𝜔(t − t0 )
𝜔 𝜔
2
= yst (𝜔(t − t0 ) − sin 𝜔(t − t0 )) (A.16)
𝜔Δt
For the third time interval,
t
F(𝜏)
y(t) = sin 𝜔(t − 𝜏)d𝜏
∫t0 M𝜔
( ( ) )
t0 + Δt
2 2(𝜏 − t )
t
0 2(𝜏 − t0 )
= yst 𝜔 sin 𝜔(t − 𝜏)d𝜏 + 2− sin 𝜔(t − 𝜏)d𝜏
∫t0 Δt ∫t0 + Δt Δt
( 2
)
t0 + Δt t
2𝜔 2
= yst (𝜏 − t0 ) sin 𝜔(t − 𝜏)d𝜏 + (Δt − (𝜏 − t0 )) sin 𝜔(t − 𝜏)d𝜏
Δt ∫t0 ∫t0 + Δt
2
Δt Δt
⎛ t0 + 2 t0 + 2 ⎞
⎜ 𝜏 sin 𝜔(t − 𝜏)d𝜏 − t0 sin 𝜔(t − 𝜏)d𝜏 ⎟
2𝜔 ⎜ ∫ t0 ∫t0 ⎟
= yst
Δt ⎜ t t

⎜ +(Δt + t ) sin 𝜔(t − 𝜏)d𝜏 − 𝜏 sin 𝜔(t − 𝜏)d𝜏 ⎟

0
∫t0 + Δt ∫t0 + Δt ⎠
2 2

⎛ [( ) ]t0 + Δt

𝜏 − t0 2
⎜ cos 𝜔(t − 𝜏) + 𝜔12 sin 𝜔(t − 𝜏) ⎟
2𝜔 ⎜ 𝜔 ⎟
[( ) t0
]
Δt ⎜⎜ ⎟
= yst t
Δt + t0 − 𝜏 ⎟
+ cos 𝜔(t − 𝜏) − 𝜔12 sin 𝜔(t − 𝜏)
⎜ 𝜔 Δt ⎟
⎝ t0 + 2 ⎠
( ) ( )
⎛ t + Δt
− t ( ) t −t ⎞
0 0 Δt

2
cos 𝜔 t − t0 − − 0 0 cos 𝜔(t − t0 )⎟
⎜ 𝜔 2 𝜔 ⎟
⎜ ( ) ⎟
1 Δt
⎜ + 2 sin 𝜔 t − t0 − ⎟
⎜ 𝜔 2 ( ) ⎟
2𝜔 ⎜ 1 Δt + t0 − t ⎟
𝜔(t 𝜔(t
Δt ⎜⎜ ⎟
= yst − sin − t ) + cos − t)
𝜔2 0
𝜔 ⎟
( )
⎜ Δt + t0 − t0 − Δt ( ) ⎟
⎜ − Δt 1
2
cos 𝜔 t − t0 − − 2 sin 𝜔(t − t) ⎟
⎜ 𝜔 2 𝜔 ⎟
⎜ ( ) ⎟
⎜ +
1
sin 𝜔 t − t −
Δt ⎟
⎝ 𝜔2 0
2 ⎠
( ( ))
2 Δt
= yst 𝜔(Δt + t0 − t) − sin 𝜔(t − t0 ) + 2 sin 𝜔 t − t0 − (A.17)
𝜔Δt 2
256 Appendix A Basic Concepts of Dynamics

Finally, for the situation after the impulse, the system would behave as a harmonic
system with the initial conditions of position and velocity which are derived from
Eq. (A.18) at time t = t0 + Δt (see Eq. (A.3)):
( ( ))
2 Δt
y(t) = yst 𝜔(Δt + t0 − t) − sin 𝜔(t − t0 ) + 2 sin 𝜔 t − t0 −
𝜔Δt 2
( )
2 Δt
y(t0 + Δt) = y0 = yst − sin 𝜔Δt + 2 sin 𝜔
𝜔Δt 2
( ( ))
2 Δt
̇ = yst
y(t) −1 − cos 𝜔(t − t0 ) + 2 cos 𝜔 t − t0 −
Δt 2
( )
2 Δt
̇ 0 + Δt) = ẏ 0 = yst
y(t −1 − cos 𝜔Δt + 2 cos 𝜔
Δt 2
ẏ 0
y = y0 cos 𝜔(t − t0 − Δt) + sin 𝜔(t − t0 − Δt) (A.18)
𝜔
As is logical, the maximum value expression of Eq. (A.18) depends only upon the
ratio of impulse duration Δt to first natural period and on the static deflection.
As an application, Figure A.6 shows the results for two systems subjected to
a triangular pulse (t0 = 0) for durations of the pulse equal to one-fourth and
five-fourths of the period. Vertical lines indicate the end of the applied impulse.
The maximum DLF for the first case is limited to 0.7 while for the second case it
reaches 1.4. These results are logical in themselves and, also, when a comparison is
done with the rectangular pulse shown in Figure A.4. Obviously, the constant pulse

ω = 14 rad/s
2.0

1.5
duration of pulse for Δt = 5T/4
1.0

0.5
DLF = y/yst

0.0
0 0.5 1 1.5 2 2.5
–0.5

–1.0

–1.5

duration of pulse for Δt = T/4


–2.0
Time (s)
DLF Δt = 5/4T DLF Δt = 0.25T

Figure A.6 Time history of a SDOF system with a natural frequency 𝜔 = 14 rad/s
(T = 0.45 s) subjected to a triangular load as a function of the duration of the impulse
(Δt = 0.25T and Δt = 4/5T).
A.1 Dynamics of Single Degree-of-Freedom Systems 257

results in a larger DLF for the same ratio of pulse duration to natural period of the
system.

A.1.2 Basics of Resonance


Dynamic effects can become much larger when periodic actions are involved, as is
the case for railway structures. When a system is excited by a periodic persistent load
whose frequency is equal to the system’s first natural frequency and damping is not
considered, deflections become infinite.
If a SDOF system, without damping and subjected to a periodic sine force, is con-
sidered, dynamic equilibrium can be written as in Eq. (A.19):
F0
M ÿ + Ky = F0 sin 𝜔f t → ÿ + 𝜔y = sin 𝜔f t (A.19)
M
The general solution to this differential equation will be the sum of the solution of
the homogenous equation (Eq. (A.3)) and a particular solution of the full equation.
It can easily be verified that a harmonic sine equation of circular frequency 𝜔f is
such a solution. Eq. (A.20) provides the solution to the problem assuming a system
initially at rest.
y = A sin 𝜔f t
ÿ = −A𝜔2f sin 𝜔f t
F0 F
sin 𝜔f t → −A𝜔2f sin 𝜔f t + A𝜔2 sin 𝜔f t = 0 sin 𝜔f t
ÿ + 𝜔y =
M M
F0
→A= ( )
M 𝜔2 − 𝜔2f
| |
| |
F0 Kymax | 1 |
→y= ( ) sin 𝜔 t → = | ( 𝜔 )2 || (A.20)
f |
M 𝜔2 − 𝜔f 2 F | |
0
|1 − 𝜔 |
f

| |
From this equation it is easy to see that if the frequency of the periodic force is close
to the natural vibration period the displacements become very large. For a system
without damping, it becomes infinity when both frequencies are equal.
Eq. (A.20) can be generalised to account for damping. The resulting expression
is given in Eq. (A.21), where 𝜉 is the damping index, that √ is the ratio between the
damping coefficient c and the critical damping value cc = 2 km = 2m𝜔:
Kymax 1
= √ (A.21)
F0 ( )2 ( ( 𝜔 )2 )2
𝜔f
2𝜔𝜉 + 1− f
𝜔

As can be seen, for damped systems, the deflection is finite but still experiences
a significant increase as the period of the action is close to the first natural period
258 Appendix A Basic Concepts of Dynamics

35.00

30.00

25.00
Kymax /F0

20.00

15.00

10.00

5.00

0.00
0 0.5 1 1.5 2 2.5 3
ωf /ω
No damping Damping ratio = 1.5%

Figure A.7 Maximum displacement as a function of the ratio 𝜔f /𝜔.

of the structure. For actual structures, the damping index varies from 0.5% for steel
structures to 2.0% for reinforced concrete structures. From Figure A.7, it can be con-
cluded that for such low damping indexes the effect of damping on the reduction of
resonance effects is modest. More important in limiting the effects of resonance in
railway structures is the fact that the number of cycles is limited, while the above
analysis considers an infinite number of cycles.

A.1.3 Solution of the Equation of Motion of a SDOF Damped System


Subjected to a Triangular Load
To solve this problem, four time intervals need to be distinguished:

– Before the impulse is applied, (t < t0 ) at which time the response of the system will
be nil
– The interval t0 ≤ t ≤ t0 + Δt/2, during which time the value of the force is
increasing
– The interval t0 + Δt/2 ≤ t ≤ t0 + Δt, during which time the value of the force is
decreasing
– After t0 + Δt when the system will behave as a freely vibrating system with initial
conditions determined by the position and the velocity of the system for t = t0 + Δt.

Before proceeding to solve the problem, it is convenient to obtain analytical


expressions for two integrals which appear repeatedly in the calculations, as well
as for their derivatives with respect to time. This is undertaken in the following
section.
A.1 Dynamics of Single Degree-of-Freedom Systems 259

A.1.3.1 Auxiliary Expressions – Integrals I 1 , I 2 , and Their Derivatives


Integral I 1 :
b
I1,a→b (t) = e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
∫a ⏟⏞⏟⏞⏟⏟⏞⏞⏞⏞⏞⏟ ⏞⏞⏞⏞⏞⏟
u dv
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
I1

b
⎡ ⎤
⎢ ⎥
⎢ −𝜉𝜔(t−𝜏) cos 𝜔d (t − 𝜏) ⎥ t
cos 𝜔d (t − 𝜏)
= ⎢e ⎥ − 𝜉𝜔e−𝜉𝜔(t−𝜏) d𝜏
⎢⏟⏞⏟⏞⏟ 𝜔 d ⎥ ∫t0 ⏟⏞⏞⏞⏟⏞⏞⏞⏟ 𝜔d
⎢ u ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟⎥ du
⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟
⎣ v ⎦a v
[ −𝜉𝜔(t−𝜏) ]b t
e 𝜔
= cos 𝜔d (t − 𝜏) − 𝜉 e−𝜉𝜔(t−𝜏) cos 𝜔d (t − 𝜏)d𝜏
𝜔d a 𝜔d ∫t0 ⏟⏞⏟⏞⏟⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟
u2 dv2
[ ]b
e−𝜉𝜔(t−𝜏)
= cos 𝜔d (t − 𝜏)
𝜔d a
b
⎛⎡ ⎤ ⎞
⎜⎢ ⎥ ( ) ⎟
𝜔 ⎜⎢ −𝜉𝜔(t−𝜏) − sin 𝜔d (t − 𝜏) ⎥ − sin 𝜔d (t − 𝜏) ⎟
b
−𝜉 ⎜⎢ e ⎥ − 𝜉𝜔e−𝜉𝜔(t−𝜏) d𝜏 ⎟
𝜔d ⎜⎢⏟⏞⏟⏞⏟ 𝜔d ⎥ ∫a ⏟⏞⏞⏞⏟⏞⏞⏞⏟ 𝜔d ⎟
⎜⎢ u2 ⏟⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏟⎥ du2 ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⎟
⎝⎣ v2 ⎦a v2 ⎠
[ −𝜉𝜔(t−𝜏) ]b
e 𝜔
= cos 𝜔d (t − 𝜏) + 𝜉 2 [e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)]ba
𝜔d a 𝜔d
( )2 b
𝜔
− 𝜉2 e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
𝜔d ∫a
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
I1

b
e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
∫a
b
⎡ ⎤
⎢ ( )⎥
e −𝜉𝜔(t−𝜏) 𝜔
=⎢ ( (

)2 ) cos 𝜔d (t − 𝜏) + 𝜉 𝜔 sin 𝜔d (t − 𝜏) ⎥
⎢ 𝜔 d
⎢ 𝜔d 1 + 𝜉 𝜔 ⎥
⎣ d ⎦a
( )
e−𝜉𝜔(t−b) cos 𝜔d (t − b) + 𝜉 𝜔𝜔 sin 𝜔d (t − b)
( d
)
−e−𝜉𝜔(t−a) cos 𝜔d (t − a) + 𝜉 𝜔𝜔 sin 𝜔d (t − a)
( )2 )
d
= I1,a→b (t) = ( (A.22)
𝜔
𝜔d 1 + 𝜉 𝜔
d
260 Appendix A Basic Concepts of Dynamics

Derivative with respect to t of integral I 1 :


( )
−𝜉𝜔e−𝜉𝜔(t−b) cos 𝜔d (t − b) + 𝜉 𝜔𝜔 sin 𝜔d (t − b)
( d )
+𝜔d e−𝜉𝜔(t−b) − sin 𝜔d (t − b) + 𝜉 𝜔𝜔 cos 𝜔d (t − b)
( d )
+𝜉𝜔e−𝜉𝜔(t−a) cos 𝜔d (t − a) + 𝜉 𝜔𝜔 sin 𝜔d (t − a)
( d )
−𝜔d e−𝜉𝜔(t−a) − sin 𝜔d (t − a) + 𝜉 𝜔𝜔 cos 𝜔d (t − a)
İ 1,a→b (t) = ( )2 )
d
( (A.23)
𝜔
𝜔d 1 + 𝜉 𝜔
d

Integral I 2 :
b
⎡ ⎤
⎢ ⎥
b
⎢ cos 𝜔 d (t − 𝜏) ⎥
I2,t0 (t) = 𝜏e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏 = ⎢𝜏e−𝜉𝜔(t−𝜏) ⎥
∫a ⏟⏞⏟⏞⏟⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟ ⎢⏟⏞⏟⏞⏟ 𝜔d ⎥
u dv ⎢ u ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟ ⎥
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⎣ v ⎦a
I2
b
cos 𝜔d (t − 𝜏)
− (e−𝜉𝜔(t−𝜏) + 𝜏𝜉𝜔e−𝜉𝜔(t−𝜏) ) d𝜏
∫a ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ 𝜔d
du ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟
v
b
⎡ ⎤
⎢ ⎥
⎢ −𝜉𝜔(t−𝜏) cos 𝜔d (t − 𝜏) ⎥ 1
b
= ⎢𝜏e ⎥ −𝜔 e−𝜉𝜔(t−𝜏) cos 𝜔d (t − 𝜏)d𝜏
⎢ ⏟⏞⏟⏞⏟ 𝜔 d ⎥ d
∫a
⎢ u ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟⎥ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
⎣ v ⎦a [ ]t
t
− e−𝜉𝜔(t−𝜏)
sin 𝜔d (t−𝜏) +𝜉 𝜔𝜔 e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
𝜔d t0 d ∫ t0
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
I1

b
𝜔
−𝜉 𝜏e−𝜉𝜔(t−𝜏) cos 𝜔d (t − 𝜏)d𝜏
𝜔d ∫a ⏟⏞⏟⏞⏟⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟
u2 dv2
b
⎡ ⎤
⎢ ⎥ [ ]b
⎢ cos 𝜔d (t − 𝜏) ⎥ 1 e−𝜉𝜔(t−𝜏)
= ⎢𝜏e−𝜉𝜔(t−𝜏) ⎥ + sin 𝜔 (t − 𝜏)
⎢⏟⏞⏟⏞⏟ 𝜔d ⎥ 𝜔d 𝜔d d
a
⎢ u ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟ ⎥
⎣ v ⎦a
b
𝜔
−𝜉 e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
𝜔2d ∫a
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
I1
b
⎛⎡ ⎤ ⎞
⎜⎢ ⎥ ⎟
𝜔 ⎜⎢ −𝜉𝜔(t−𝜏) − sin 𝜔d (t − 𝜏) ⎥ − sin 𝜔d (t − 𝜏) ⎟
b
−𝜉 𝜏e −𝜉𝜔(t−𝜏)
𝜏𝜉𝜔e −𝜉𝜔(t−𝜏)
𝜔d ⎜⎜⎢⎢⏟⏞⏟⏞⏟ ⎥ ⎟
− (e + ) d𝜏
𝜔d ⎥ ∫a ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ 𝜔d ⎟
⎜⎢ u 2
⏟⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏟ ⎥ du2 ⏟⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏟ ⎟
⎝⎣ v2 ⎦a v2 ⎠
A.1 Dynamics of Single Degree-of-Freedom Systems 261

b
⎡ ⎤
⎢ ⎥ [ ]b
⎢ −𝜉𝜔(t−𝜏) cos 𝜔d (t − 𝜏) ⎥ 1 e−𝜉𝜔(t−𝜏) 𝜔
= ⎢𝜏e ⎥ + sin 𝜔d (t − 𝜏) − 𝜉 2 I1,a→b
⎢⏟⏞⏟⏞⏟ 𝜔d ⎥ 𝜔d 𝜔d a 𝜔d
⎢ u ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟⎥
⎣ v ⎦a
b
𝜔 𝜔
+𝜉 [𝜏e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)]ba − 𝜉 2 e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
𝜔2d 𝜔d ∫ a
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
I1,a→b
( )2 b
𝜔
− 𝜉2 𝜏e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏 →
𝜔d ∫a
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
I2
b
𝜏e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏
∫a
[ ( ( ) )]b
e−𝜉𝜔(t−𝜏)
𝜔d
𝜏 cos 𝜔d (t − 𝜏) + 1
𝜔d
+ 𝜏𝜉 𝜔𝜔 sin 𝜔d (t − 𝜏) − 2𝜉 𝜔𝜔2 I1,a→b
d a
( )2 )
d
= (
1 + 𝜉 2 𝜔𝜔
d
( ( ) )
⎛ e −𝜉𝜔(t−b)
b cos 𝜔d (t − b) + 𝜔 + b𝜉 𝜔𝜔 sin 𝜔d (t − b)
1 ⎞
⎜ ( ( d d ) ) ⎟
⎜ −e −𝜉𝜔(t−a)
a cos 𝜔d (t − a) + 𝜔 + a𝜉 𝜔 sin 𝜔d (t − a) − 2𝜉 𝜔 I1,a→b ⎟⎠
1 𝜔 𝜔

( ( )2 )
d d d
I2,a→b (t) = (A.24)
𝜔d 1 + 𝜉 2 𝜔𝜔
d

Derivative with respect to t of integral I 2 :


( ( ) )
⎛ −𝜉𝜔e−𝜉𝜔(t−b) b cos 𝜔 (t − b) + 1 + b𝜉 𝜔 sin 𝜔 (t − b) ⎞
⎜ d 𝜔d 𝜔d d ⎟
⎜ ( ( ) )⎟
1 𝜔
⎜ +𝜔d e −𝜉𝜔(t−b) −b sin 𝜔d (t − b) + 𝜔 + b𝜉 𝜔 cos 𝜔d (t − b) ⎟
⎜ ( (
d
)
d
) ⎟
⎜ +𝜉𝜔e−𝜉𝜔(t−a) a cos 𝜔 (t − a) + 1 + a𝜉 𝜔 sin 𝜔 (t − a) ⎟
⎜ d 𝜔d 𝜔d d ⎟
⎜ ( ( ) )⎟
1 𝜔
⎜ −𝜔d e −𝜉𝜔(t−a)
−a sin 𝜔d (t − a) + 𝜔 + a𝜉 𝜔 cos 𝜔d (t − a) ⎟
⎜ d d

⎜ −2𝜉 𝜔 dI1,a→b

⎝ 𝜔d dt ⎠
İ 2,a→b (t) = ( ( )2 )
𝜔d 1 + 𝜉 2 𝜔𝜔
d

(A.25)

A.1.3.2 Solution for the damped SDOF System Subjected to a Triangular Load
For the second time interval, t0 ≤ t ≤ t0 + Δt/2:
t t
F(𝜏) F
y(t) = e−𝜉𝜔(t−𝜏) sin 𝜔d (t − 𝜏)d𝜏 = max e−𝜉𝜔(t−𝜏) f (𝜏) sin 𝜔d (t − 𝜏)d𝜏
∫0 M𝜔d M𝜔d ∫0
Δt
if t0 ≤ t ≤ t0 +
2
Fmax t −𝜉𝜔(t−𝜏) 2(𝜏 − t0 )
y(t) = e sin 𝜔d (t − 𝜏)d𝜏
M𝜔d ∫t0 Δt
262 Appendix A Basic Concepts of Dynamics

( t t )
Fmax 2
= e −𝜉𝜔(t−𝜏)
𝜏 sin 𝜔d (t − 𝜏)d𝜏 − t0 e −𝜉𝜔(t−𝜏)
sin 𝜔d (t − 𝜏)d𝜏
M𝜔d Δt ∫t0 ∫t0
F 2
y(t) = max (I (t) − t0 I1,t0 →t (t)) (A.26)
M𝜔d Δt 2,t0 →t
For the third time interval, t0 + Δt/2 ≤ t ≤ t0 + Δt
Δt
if t0 + ≤ t ≤ t0 + Δt
2
Δt
Fmax t0 + 2 −𝜉𝜔(t−𝜏) (𝜏 − t0 )
y(t) = 2 e sin 𝜔d (t − 𝜏)d𝜏
M𝜔d ∫t0 Δt
( )
Fmax t (𝜏 − t0 )
+2 e −𝜉𝜔(t−𝜏)
1− sin 𝜔d (t − 𝜏)d𝜏
M𝜔d ∫t0 + Δt Δt
2

F ( )
2
y(t) = max I2,t0 →t0 + Δt (t) − t0 I1,t0 →t0 + Δt (t)
M𝜔d Δt 2 2

t ( )
Fmax (𝜏 − t0 )
+2 e −𝜉𝜔(t−𝜏)
1− sin 𝜔d (t − 𝜏)d𝜏
M𝜔d ∫t0 + Δt Δt
2

F ( )
2
= max I2,t0 →t0 + Δt (t) − t0 I1,t0 →t0 + Δt (t)
M𝜔d Δt 2 2
(( ) ( ) )
Fmax t0 1
+2 1+ I Δt (t) − I Δt (t)
M𝜔d Δt 1,t0 + 2 →t Δt 2,t0 + 2 →t
F ( ( )
1
y(t) = 2 max I2,t0 →t0 + Δt (t) − I2,t0 + Δt →t (t) − t0 I1,t0 →t0 + Δt (t)
M𝜔d Δt 2 2 2
( ) )
t0
+ 1+ I Δt (t) (A.27)
Δt 1,t0 + 2 →t
Finally, for the fourth time interval (t ≥ t0 + Δt), the system will be in free vibration
with the initial velocity and position attained at time t0 + Δt:
y0 = y(t0 + Δt)
F ( ( )
1
= 2 max I2,t0→ t0 + Δt (t0 + Δt) − I2,t0 + Δt →t0 +Δt (t0 + Δt) − t0 I1,t0→ t0 + Δt (t0 + Δt)
M𝜔d Δt 2 2 2
( ) )
t
+ 1 + 0 I1,t0 + Δt →t0 +Δt (t0 + Δt)
Δt 2

Fmax
ẏ 0 = 2
M𝜔d
( ( )
1 ̇
× I 2,2,t0→ t0 + Δt (t0 + Δt) − İ 2,t0 + Δt →t0 +Δt (t0 + Δt) − t0 İ 1,2,t0→ t0 + Δt (t0 + Δt)
(Δt ) )
2 2 2

t0
+ 1+ İ Δt (t + Δt)
Δt 1,t0 + 2 →t0 +Δt 0
( )
ẏ 0 + 𝜉𝜔y0
y(t) = e −𝜉𝜔(t−t0 −Δt)
y0 cos 𝜔d (t − t0 − Δt) + sin 𝜔d (t − t0 − Δt) (A.28)
𝜔d

Reference

1 Biggs, J. (1964). Introduction to Structural Dynamics. New York: McGraw-Hill


Book Company.
263

Appendix B

Singular Bridges for High-Speed Railway Lines


José Romo

B.1 Germany
A number of special bridges built in Germany are listed below. The first four
(Gemünden, Veitshöchheim, Pfieffetal, and Nantenbach) correspond to the con-
struction of the first German high-speed railways built mainly in the second part
of the 1980s and the first part of the 1990s. The rest are special bridges built after
German reunification in the first two decades of the 21st century.
The German bridges built in the first stage were designed so that any deck could be
replaced in the shortest period of time. Therefore, German bridges of that period are
generally isostatic, and in the case of continuous bridges, their length was limited
(to 400 m) to allow for a hypothetical rapid replacement.
Bridges built in Germany in the second decade of the 21st century have been
developed according to the guidelines of the DB Netze Railway Bridge Design Guide
2008. With the implementation of this guide, the condition that the decks should be
quickly replaceable has been removed and semi-integral bridges have been designed,
generally without track expansion joints and structure joints every 100–120 m or
so. Examples of such solutions are the Unstruttal bridge and the Gänsebachtal
viaduct.

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
264 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.1 Gemünden Bridge


B.1.1.1 Data Summary
Owner Deutsche Bahn AG
Place Gemünden am Main, Germany
HSR Line Hannover – Würzburg
Designer LAP – Leonhardt, Andrä und Partner
Contractor Max Streicher GmbH & Co. KG
Mayreder, Kraus & Co. Baugesellschaft mbH
Main span 135.0 m
Deck Width 14.3 m
Length 793.5 m
Deck type Box girder
Material Concrete
Typology Frame
Start construction 1982
Completion 1984

B.1.1.2 Description
This bridge has a total length of 793.5 m and is divided into three continuous sections
of 164.0 – 299.0 – 330.5 so that it can be replaced in parts, with typical spans of
55.0 m. The braking is independently resisted in each sector. The lateral spans are
anchored to the abutments, which are the elements that take the braking loads. In
the intermediate section, V-shaped piers, the braking loads are transmitted through
the hinge located at the base of the V-shaped pier. The deck is 4.50 m in depth in
general, increasing to 6.50 m in the area of the V-shaped piers. The main span has a
length of 135 m. The track has only two expansion joints at the ends of the V-deck
section.

Further Reading
Leonhardt, F. (1986). Mainbrücke Gemünden – Eisenbahnbrücke aus Spannbeton mit
135 m Spannweite. Beton- und Stahlbetonbau 81 (1): 1–8.
Zellner, W. and Saul, R. (1991). Long span bridges of the new railroad lines in Germany.
IABSE Report 64 (1991).
https://structurae.net/en/structures/gemunden-high-speed-rail-viaduct.
B.1 Germany 265

Figure B.1 Gemünden Bridge (Courtesy of Deutsche Bahn AG).

Figure B.2 Gemünden Bridge (Courtesy of Deutsche Bahn AG).


266 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.2 Veitshöchheim Bridge


B.1.2.1 Data Summary
Owner Deutsche Bahn AG
Place Veitshöchheim, Germany
HSR Line Hannover – Würzburg
Designer ILF Consulting Engineers/Leonhardt,
Andrä und Partner
Contractor Strabag AG; Walter Thosti Boswau AG
Main span 162.0 m
Deck Width 14.0 m
Length 1.280 m
Deck type Box girder
Material Concrete
Typology Arch + Beam
Start construction 1985
Completion 1987

B.1.2.2 Description
The bridge over the river Main in Veitshöchheim has a total length of 1264 m. It is
divided into five separate parts of 237.0, 369.5, 214.0, 160.5, and 281.6 m. The track
has rail joints in the two abutments and in the pier between the 214.0 and 160.5 m
long sections. The deck is anchored longitudinally at the arch keystone and at the
three piers located at the transition between the 160.5 and 281.6 m sections.
The 237.0, 369.5, and 214.0 deck sections are connected by shock transmitters
located in the transition pile sections that allow slow movements but are locked to
transmit the braking loads to the arch keystone.

Further Reading
Zellner, W. and Saul, R. (1991). Veitshöchheim viaduct: a concrete arch bridge with
162 m main span. IABSE Report 64.
Zellner, W. and Saul, R. (1991). Long span bridges of the new railroad lines in Germany.
IABSE Report 64.
https://structurae.net/en/structures/veitshochheim-viaduct.
B.1 Germany 267

Figure B.3 Veitshöchheim Bridge (Courtesy of Deutsche Bahn AG).

Figure B.4 Veitshöchheim Bridge (Source: Störfix - Eigenes Werk, CC BY-SA 3.0,
https://commons.wikimedia.org/w/index.php?curid=916591).
268 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.3 Pfieffetal Bridge


B.1.3.1 Data Summary
Owner Deutsche Bahn AG
Place Adelshausen, Germany
HSR Line Hannover – Würzburg
Designer Harries + Kinkel; Werner Sobek
Main span 116 m
Deck Width 14.3 m
Length 812 m
Deck type Box-girder
Material Concrete
Typology Beam
Start construction 1987
Completion 1989

B.1.3.2 Description
This structure has a total length of 17 × 58 = 986 m. The deck is made up of 17 spans
of 55.75 m span and 5.30 m depth. The spans are isostatic so that they can be replaced
quickly and independently. Each deck span is connected to its adjacent spans by
pre-stressed cables to transmit the braking loads. Due to the considerable height
above the valley (approx. 95 m), a central A-pier has been provided to transmit the
braking loads to the ground. The shafts of the inclined piers are curved to eliminate
bending under their own weight. The track only has two expansion joints at the
abutments. The track has expansion joints only in the abutments.

Further Reading
Harries, H., Kinkel, H., and Petri, H. (1987). Die Rombachtalbrücke. Beton- und
Stahlbetonbau 82 (7): 179–185.
Zellner, W. and Saul, R. (1991). Long span bridges of the new railroads lines in
Germany. IABSE Report 64.
B.1 Germany 269

Figure B.5 Pfieffetal Bridge (Courtesy of Wolfgang Pehlemann).

Figure B.6 Pfieffetal Bridge (Courtesy of Wolfgang Pehlemann).


270 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.4 Nantenbach Bridge


B.1.4.1 Data Summary
Owner Deutsche Bahn AG
Place Neuendorf, Germany
HSR Line Hannover – Würzburg
Designer LAP – Leonhardt, Andrä und Partner
Main span 208.0 m
Deck Width 14.3 m
Length 694.5 m
Deck type Box girder
Material Concrete
Typology Truss
Start construction 1991
Completion 1993

B.1.4.2 Description
The first section of the bridge is 320.10 m long and has spans of 50 m. The deck of
this part is a hollow box concrete section. The second sector is a 374.4 m long con-
tinuous mixed truss with spans of 83.2, 208.0, and 83.2 m. The deck depth is 8.50 in
the central area and 16.50 in the support sections of the main piers, giving a slender-
ness of 1/24 and 1/13, respectively. The truss has a double composite action, with a
bottom slab in the lower part of the support sections in the main piers.

Further Reading
Zellner, W. and Saul, R. (1991). Railway bridge with double composite action across
river Main. IABSE Report 64.
Zellner, W. and Saul, R. (1991). Long span bridges of the new railroad lines in Germany.
IABSE Report 64.
https://structurae.net/en/structures/railroad-bridge-at-nantenbach.
B.1 Germany 271

Figure B.7 Nantenbach Bridge (Courtesy of Deutsche Bahn AG).

Figure B.8 Nantenbach Bridge (Courtesy of Deutsche Bahn AG).


272 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.5 Unstruttal Bridge


B.1.5.1 Data Summary
Owner Deutsche Bahn AG
Place Karsdorf, Sachsen-Anhalt, Germany
HSR line Erfurt/Leipzig – Halle
Designer sbp schlaich bergermann partner
Contractor Alpine Bau Deutschland GmbH;
Berger Bau
Arch spans 4 × 108 m
Deck width 13.95–15.93 m
Length 2668 m
Deck type Box girder
Material Concrete
Typology Arch – Beam
Start construction 2007
Completion 2012

B.1.5.2 Description
This viaduct is made up of four sections of 580 m long, each consisting of 10 con-
tinuous spans. The spans have a typical length of 58 m, and the fixed points of each
section are materialised in the keystones of the arches that are located in the centre
of the span. The piers are very slender and all are connected monolithically at their
head to the deck. This results in a structure in which the only bearing devices of the
deck are in the abutments. There are track expansion joints at the same position as
the structural expansion joints.

Further Reading
Rimane, T. (2011). Der Bau der Eisenbahnüberführung Unstruttalbrücke. In: Deutscher
Bautechnik-Tag 11-13. Mai 2011 (ed. DBV), 69–70. Berlin: Tagungsband.
Schlaich, M. (2012). Integral Railway Bridges in Germany. 22nd Dresdener
Brückenbausymposium. Dresden.
https://structurae.net/en/structures/unstrut-viaduct.
B.1 Germany 273

Figure B.9 Unstruttal Bridge (Courtesy of Deutsche Bahn AG).

Figure B.10 Unstruttal Bridge (Courtesy of Deutsche Bahn AG).


274 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.6 Gänsebachtal Viaduct


B.1.6.1 Data Summary
Owner Deutsche Bahn AG
Place Buttstädt, Germany
HSR Line Erfurt/Leipzig – Halle
Designer sbp schlaich bergermann partner
Contractor Adam Hörnig Baugesellschaft GmbH & Co.;
Stutz GmbH Tief- und Straßenbau
Typical span 24.50
Deck Width 13.83
Length 1012 m
Deck type Twin-webbed T monolithic beam
Material Concrete
Typology Beam
Start construction 2007
Completion 2012

B.1.6.2 Description
This viaduct consists of 10 sections of 112 m each separated by structural expansion
joints. Each segment has five spans with a rigid double pier which is the fixed point of
each segment. All piers are monolithically embedded at their heads in the deck, and
therefore there are only bearings on the abutments. The track is continuous along
the entire length of the bridge and there are no track expansion joints, not even at the
abutments.

Further Reading
Schenkel, M., Goldack, A., and Schlaich, J. (2010). Die Gänsebachtalbrücke, eine
integrale Talbrücke der DB AG auf der Neubaustrecke Erfurt-Leipzig/Halle. Beton-
und Stahlbetonbau 105 (9): 590–598. https://doi.org/10.1002/best.201000034.
Schlaich, M. (2012). Integral Railway Bridges in Germany. 22nd Dresdener
Brückenbausymposium. Dresden.
https://structurae.net/en/structures/gansebach-viaduct.
B.1 Germany 275

Figure B.11 Gänsebachtal Viaduct (Courtesy of Störfix)


https://commons.wikimedia.org/wiki/File:G%C3%A4nsebachtalbr%C3%BCcke-2018-07a.

Figure B.12 Gänsebachtal Viaduct (Courtesy of Deutsche Bahn AG).


276 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.7 Hämerten Bridge


B.1.7.1 Data Summary
Owner Deutsche Bahn AG
Place Hämerten, Germany
HSR Line Hannover – Berlin
Designer Ingenieurbüro Grassl GmbH
Contractor Philipp Holzmann AG
Main span 105.8 m
Deck Width 14.3 m
Length 810 m
Deck type Steel grid + box girder
Material Steel (truss) + concrete
Typology Truss
Start construction 2006
Completion 2013

B.1.7.2 Description
The superstructure of the 18-span bridge consists of a western 12-span bridge made
of prestressed concrete, the 3-span current bridge made of steel, and the eastern
3-span foreshore bridge made of prestressed concrete.
The total length is 809.87 m. It consists of one section with 12 spans of 34.07 m
– 34.25 m – 34.16 m – 34.21 m – 34.20 m – 34.21 m – 34.20 m – 41.11 m – 45.18 m
– 45.06 m – 43.62 m, of the truss structure section with spans of 66.99 m – 105.77 m
– 66.9 m, and the third sector with spans of 40.66 m – 40.55 m – 40.52 m.
Horizontal braking forces are transferred to the east abutment and the pier
between the steel bridge and the west flood bridge. The track expansion joints are
located at the abutment west and above the pier between the steel bridge and the
flood bridge east.

Further Reading
Braun, M. (2013). Die Eisenbahnbrücke bei Hämerten in Sachsen-Anhalt. Bautechnik
90 (2): 113–119. https://doi.org/10.1002/bate.201200039.
Heinisch, R., Siegmann, J., Stuchly, H., and Witt, P. (2001). Ingenieurbauwerke.
Hestra-Verlag, Darmstadt (Germany): ETR - Eisenbahntechnische Rundschau.
https://structurae.net/en/structures/hamerten-railroad-bridges.
B.1 Germany 277

Figure B.13 Hämerten Bridge (Courtesy of Mefellingen).

Figure B.14 Hämerten Bridge (Courtesy of Störfix).


278 Appendix B Singular Bridges for High-Speed Railway Lines

B.1.8 Filstal Bridge


B.1.8.1 Data Summary
Owner Deutsche Bahn AG
Place Mühlhausen im Täle, Germany
HSR Line Wendlingen – Ulm
Designer ILF Consulting Eng./LAP – Leonhardt,
Andrä und Partner
Main span 162.0 m
Deck Width 14.0 m
Length 485 m
Deck type Box girder
Material Concrete
Typology Frame
Start construction 2014
Completion 2022 foreseen

B.1.8.2 Description
The bridge connects the Boßlertunnel and the Steinbühltunnel. The structure con-
sists of two single-track, parallel prestressed concrete bridges, 485 and 472 m long,
respectively. The spans of the western structure are 44, 95, 150, 93, 58, 45 m and of
the eastern structure 44, 95, 150, 88, 50, 45 m. The two bridge superstructures are
designed as single-cell box girder cross-sections prestressed of 9.20 width. They are
monolithically connected to the bridge piers. On each abutment, there is an expan-
sion joint and the deck is resting in bearings. The frame construction is referred to
as semi-integral construction.

Further Reading
Angelmaier, V. (2004). Neubaustrecke Wendlingen – Ulm. Die Filstalbrücke als Teil des
"Albaufstiegs". Umrisse 4 (2): 13–17.
Schlaich, M. (2012): Integral Railway Bridges in Germany. 22nd Dresdener
Brückenbausymposium, Dresden.
B.1 Germany 279

Figure B.15 Filstal Bridge (Courtesy of Deutsche Bahn AG).

Figure B.16 Filstal Bridge (Courtesy of Deutsche Bahn AG, photo: Arnim Kilgus).
B.2 France 281

B.2 France
Unlike the first generation of German bridges, French bridges were not designed so
that their decks could be replaced quickly if necessary. This has led to bridges on
high-speed lines in France having continuous decks. The use of continuous decks
means that for long viaducts it is necessary to have track expansion joints.
In these cases, French bridges use a so-called neutral or inert span. This is an
independent span, simply supported, and usually located in the centre of the bridge,
which allows the effective expansion length of the bridge to be divided in two. Thus,
the two long side decks are usually fixed to the respective abutments and two rail
joints are placed using the two expansion joints of the central neutral span.
In the case of very long viaducts, instead of anchoring the decks to the abutments,
they have been anchored to intermediate piers so that the fixed points are approx-
imately in the centre of each section. In these cases, track expansion joints are
arranged to coincide with the expansion joints of the structure at the abutments.
In addition to the aspects relating to the long viaducts, French bridges have the
particularity of the joint work between engineers and architects on the unique
bridges, with a result that is not always fortunate.
282 Appendix B Singular Bridges for High-Speed Railway Lines

B.2.1 Garde-Adhémar Viaduct


B.2.1.1 Data Summary
Owner SNCF
Place La Garde-Adhémar, France
HSR Line Lyon – Marseilles
Architect Marc Mimram
Structural Engineer Bureau d’Etudes Greisch
Main span 2 × 110.3 m
Deck Width 20.0 m
Length 324.6 m
Deck type Steel grid
Material Steel
Typology Bowstring arch
Start construction 1996
Completion 1998

B.2.1.2 Description
The flagship structure of the Lyon-Mediterranean high-speed line, it is 324.60 m long
consisting of four spans, including two 52 m approach spans and two main bowstring
spans of 110.30 m. A central arch joins the two bowstring spans and slightly increases
the overall rigidity. The steel arches support a concrete deck.
The two arches have a height above the deck of 15 m, giving a rise/span ratio of
1/7.7, while the maximum height of the central arched part has its highest point at
a distance of 24 m above the deck. The central pier supporting the two arches is the
fixed point of the bridge and has been designed to withstand the braking loads.

Further Reading
Mimram, M. Le viaduc de La Garde-Adhémar. In: Formes et Structures, n. 127.
Ramondenc, P., Plu, B., Pourcelot, M., and Hoorpah, W. (1992). La Garde-Adhémar
Bolletin 19. Pont métalliques OUTA.
https://structurae.net/en/structures/garde-adhemar-viaduct.
B.2 France 283

Figure B.17 La Garde-Adhémar Viaduct (Courtesy of Greisch).

Figure B.18 La Garde Adhémar Viaduct (Source: https://fr.wikipedia.org/wiki/Viaduc_de_


La_Garde-Adh%C3%A9mar#/media/Fichier:Bow-String_LGV-Med).
284 Appendix B Singular Bridges for High-Speed Railway Lines

B.2.2 Avignon Viaducts


B.2.2.1 Data Summary
Owner SNCF
Place Roquemaure, France
HSR Line Paris – Marseilles
Architect Jean François Blassel
Structural Engineer Michel Virlogeux; RFR; Setec TPI
Contractor Bouygues Travaux Publics
Main span 100.0 m
Deck Width 10.0 m
Length 1500 m
Deck type Box girder
Material Concrete
Typology Continuous beam
Start construction 1996
Completion 1999

B.2.2.2 Description
This bridge is divided into three 450 – 650 – 400 m sections. The first two sections
have 100 m spans and the last one has 50 m spans. The connection between the first
two sections is made by means of a half corbel joint. The first section is fixed to
the two piers closest to the half corbel joint, while the central section of the deck
is joined to two piers centred on the section of considerable height. The track has
expansion joints coinciding with the expansion joints between sections and with
the abutments.

Further Reading
Blassel, J.F., Desvignes, M., and Virlogeux, M. Les viaducs sur le Rhône à Avignon. In:
Formes et Structures, n. 127.
Chatelard, P., Martin, O., Roujon, M., and Sayn, P. (1998). Lot 2H - Les viaducs
d’Avignon. In: Travaux. n. 742.
B.2 France 285

Figure B.19 Avignon Viaducts (Courtesy of Vinci Construction).

Figure B.20 Avignon Viaducts, bearing (Courtesy of Vinci Construction).


286 Appendix B Singular Bridges for High-Speed Railway Lines

B.2.3 Mornas Viaduct


B.2.3.1 Data Summary
Owner SNCF
Place Mornas, France
HSR Line Paris – Lyon - Marseilles
Architect Jean Pierre Duval
Structural Engineer Bureau d’etudes Greisch
Contractor Campenon Bernard; EMCC; Spie Batignolles TP
Main span 121.40 m
Deck Width 14.3 m
Length 887 m
Deck type Box girder
Material Steel (arch) + concrete
Typology Bowstring + beam
Start construction 1997
Completion 1999

B.2.3.2 Description
The access spans to the bow-string are composite concrete slabs with double “T”
steel beams with spans ranging from 45 to 60 metres. One of them is the inert span
that houses the track expansion devices. The inert span is right next to the bow-string
section. In order to improve the performance of the system, the overlapping arches
are connected, in addition to a system of general bars, by means of solid zones in the
keystones and at the start of the arches.

Further Reading
Ramondenc, P. and Bousquet, C. (2004). The main bridges of the high speed line HSL
Méditerranée. In: Presented at: fib Symposium, Avignon.
Vallée, P., Roujon, M., and Sayn, P. (1998). Lot 2C - Le viaduc de Mornas. In: Travaux,
n. 742.
B.2 France 287

Figure B.21 Mornas Viaduct (Courtesy of Bureau Greisch).

Figure B.22 Mornas Viaduct (Courtesy of Véronique Pagnier).


288 Appendix B Singular Bridges for High-Speed Railway Lines

B.2.4 Savoureuse Viaduct


B.2.4.1 Data Summary
Owner SNCF
Place Bermont, France
HSR Line Rhin-Rhône
Architect Wilkinson Eyre
Design Egis Ouvrages d’Art
Contractor Eiffel Construction Metallique
Main span 121.40 m
Deck Width 14.3 m
Length 792 m
Deck type Ladder girder
Material Composite: steel + concrete
Typology Beam
Start construction 2007
Completion 2011

B.2.4.2 Description
The structure is composed of a succession of mixed-frame portals of a constant and
deliberately short length (66 m) to reduce the depth of the deck. Due to the curvature
of the route and its length, it was not possible to use the usual expansion joints to
adopt a continuous structure on supports. Therefore, the frames are independent of
each other. They consist of a short span of 21 m embedded in the supports, and an
isostatic span of 45 m resting on the previous ones by means of spherical supports
that allow longitudinal displacements and rotations. There are no track expansion
joints.

Further Reading
de la Savoureuse, V. (2005). France. In: The Architectural Review.
Yazbeck, N. (2010). LGV Rhin-Rhône, Branche Est - Les études du viaduc de la
Savoureuse. In: Travaux, n. 870.
B.2 France 289

Figure B.23 Savoureuse Viaduct (Courtesy of Wilkinson Eyre).

Figure B.24 Savoureuse Viaduct (Courtesy of Wilkinson Eyre).


B.3 Spain 291

B.3 Spain
Spanish high-speed long bridges are characterised by the use of continuous solutions
with a fixed point usually at an abutment or at one or more intermediate points in
the case of extra-long bridges.
This generally leads to the need to place a track expansion joint in the movable
abutment or two track expansion joints in abutments when the fixed point is in the
centre of the bridge.
The Spanish orography has given rise to a series of long-span bridges in which
structural rigour has been combined with visual appraisal and landscaping consid-
erations.
292 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.1 Osera Bridge


B.3.1.1 Data Summary
Owner ADIF
Place Zaragoza, Spain
HSR Line Madrid – Barcelona – French Border
Designer Carlos Fernández Casado CFC
Contractor Vías y Construcciones
Main span 120 m
Deck Width 15 m
Length 546 m
Deck type Concrete U section
Material Concrete
Typology Continuous beam
Start construction 2000
Completion 2002

B.3.1.2 Description
The bridge has a total length of 546 m with a span distribution of 18 + 6 × 24 + 60 +
120 + 2 × 60 + 42 m. It is a novel deck in which an attempt has been made to adapt
the concept of the metal trusses used in large railway bridges to the structural and
constructive domain of prestressed concrete bridges.
The deck is a box girder with lightened webs with circular holes and joined at the
top with ribs. Structurally, it is a vierendeel beam. It has a total depth of 9.15 m. The
webs are lightened with circular holes of 3.80 m. in diameter every 6.00 m. The deck
has a set of transverse beams cross beams with a circular profile depth separated by
3.0 m. with a trapezoidal section with a variable width from 0.50 to 0.60 m.

Further Reading
Manterola, J. and Martínez-Cutillas, A. (2003). The Ebro River Bridge, a new concrete
bridge for railways. In: Presented at: IABSE Symposium: Structures for High-Speed
Railway Transportation, Antwerp, Belgium.
Manterola, J., Martínez-Cutillas, A., and López, J.L. (2003). High-speed railway bridge
over the Ebro River, Spain. Structural Engineering International 13 (3): 174–176.
B.3 Spain 293

Figure B.25 Osera Bridge, interior view (Courtesy of Carlos Fernández Casado CFC & ADIF).

Figure B.26 Osera Bridge (Courtesy of Carlos Fernández Casado CFC & ADIF).
294 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.2 Llinars Del Vallès Viaduct


B.3.2.1 Data Summary
Owner ADIF
Place Llinar del Vallés, Spain
HSR Line Madrid – Barcelona – French border
Designer PEDELTA
Contractor Contractora Hispánica
Main span 75 m
Deck Width 14 m
Length 574 m
Deck type Continuous composite deck
Material Steel and Concrete
Typology Rigid inverted frame
Start construction 2003
Completion 2005

B.3.2.2 Description
The 574 m long bridge is comprised of two parts: a 307 m long steel–concrete struc-
ture crossing the busy highway AP-7 and a continuous prestressed concrete box
girder viaduct over the Mogent River with a longest span of 48 m. The steel bridge is
a continuous structure with five spans measuring 45 + 71 + 75 + 71 + 45 m, respec-
tively. The location of the piers was controlled by the highly skewed angle of the
highway crossing and by the launching process used to erect the bridge. The bridge
is designed as a king post truss with curved diagonals, which supports a composite
steel–concrete deck. The 14 m wide deck accommodates two ballasted tracks, for a
total bridge width of 17 m. The composite deck rests on 1 m deep transverse steel
plate I-beams positioned 3.55 m apart. These beams are connected to the bottom
chord, which is a 1.6 m wide box section with depth varying between 3.5 and 6 m.

Further Reading
Sobrino, J.A. (2010). Two steel bridges for the high-speed railway line
Madrid-Barcelona-French border. In: Presented at: IABSE Symposium: Improving
Infrastructure Worldwide, Weimar, Germany.
Sobrino, J.A. (2010). Realizaciones Españolas. Diez años de ingeniería estructural.
Asociación Científico-técnica del Hormigón Estructural (Cinter Divulgación
Técnica).
B.3 Spain 295

Figure B.27 Llinars des Vallès Viaduct (Courtesy of PEDELTA).

Figure B.28 Llinars del Vallès Viaduct (Courtesy of PEDELTA, photo: Alex Bo).
296 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.3 Salto Del Carnero Railway Bridge, Saragossa


B.3.3.1 Data Summary
Owner ADIF
Place Zaragoza, Spain
HSR Line Madrid – Barcelona – French border
Designer CESMA Ingenieros
Contractor Ferrovial Agromán – FCC
Main span 37 m
Deck Width 8.20 m
Length 124 m
Deck type Continuous composite deck
Material Steel and Concrete
Typology Rigid inverted frame
Start construction 2004
Completion 2006

B.3.3.2 Description
Situated only one hundred metres from Delicias Station, the overpass spans eight
railway tracks, including the Madrid-Barcelona high-speed railway as well as
regional and commuter train tracks. The deck rigid stays consist of welded steel
box sections. The pier shafts are rigidly connected to the deck. The reduction of the
height of the cross section towards the top of the pier contributes to reducing
the bending moments in these members. The bridge only has support bearings on
the abutments. Track has no expansion joint.

Further Reading
Tanner, P. and Bellod, J.L. (2006). Salto del Carnero Railway Bridge Saragossa Spain.
IABSE Structural Engineering International Journal 16 (3): 200–203.
Tanner, P. and Bellod, J.L. (2011). Concepción y proyecto del puente ferroviario
extradosado Salto del Carnero de Zaragoza- Delicias. Hormigón y Acero 62 (259).
B.3 Spain 297

Figure B.29 Salto del Carnero railway bridge (Courtesy of CESMA Ingenieros).

Figure B.30 Salto del Carnero railway bridge, view from below (Courtesy of CESMA
Ingenieros).
298 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.4 Viaduct Over AP7 Riudellots de la Selva


B.3.4.1 Data Summary
Owner ADIF
Place Riudellots de la Selva, Spain
HSR Line Madrid – Barcelona – French border
Designer FHECOR
Contractor ACCIONA
Main span 56 m
Deck Width 14 m
Length 112 m
Deck type Continuous composite deck
Material Steel
Typology Steel Truss
Start construction 2004
Completion 2006

B.3.4.2 Description
High-speed railway line between Barcelona and the French border crosses the AP-7
highway in Riudellots de la Selva (Girona). The structure has two spans of 53.00 m
and a big skew (34∘ ). A truss structure with variable depth has been used due to a
wide range of facts such as high visibility from the highway, strong stiffness require-
ments, and span length. This structural type combines stiffness and lightness.
The structure has two lateral beam trusses with a parabolic shape. Beams are
entirely closed in central area to have enough bending resistance and to transmit
the shear forces through the central support. The deck has transversal beams con-
necting the trusses without any diaphragms between piers, which creates a clean
bottom view from the highway. A concrete slab was constructed over transversal
beams. This slab directly resists railways loads and gives transversal stiffness to the
structure. The bridge is seen by a thousand people daily, and the design aim is the
combination of robustness and aesthetic appearance.

Further Reading
Romo, J. (2010). Riudellots High Speed Line Bridge. In: Presented at IABSE Symposium:
Large Structures and Infrastructures for Environmentally Constrained and Urbanised
Areas, Venice, Italy.
Romo, J. (2011). Puente de la línea de alta velocidad Barcelona-Frontera francesa sobre
la AP-7 en Riudellots de la Selva. Revista Hormigón y Acero 62 (261).
B.3 Spain 299

Figure B.31 Viaduct over AP7 Riudellots de la Selva, aerial view (Courtesy of Talyr).

Figure B.32 Viaduct over AP7 Riudellots de la Selva (Courtesy of ADIF).


300 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.5 Contreras Bridge


B.3.5.1 Data Summary
Owner ADIF
Place Contreras Reservoir, Spain
HSR Line Madrid – Levante
Designer Carlos Fernández Casado CFC
Contractor AZVI-Contractora San José
Main span 261 m
Deck Width 14.20 m
Length 587 m
Deck type Continuous box girder
Material Concrete
Typology Arch
Start construction 2007
Completion 2009

B.3.5.2 Description
The arch, with a span of 261 m, was divided into six parts by vertical columns.
The arch follows a polygonal guideline. The antifunicular of the arch is perfect in
this way, reducing the deflections that would exist in the area between the vertical
columns if the arch were perfectly curved.
It is a reinforced concrete arch bridge with a continuous prestressed concrete box
deck and two access viaducts. The span of the arch is 261 m and a rise at the centre
is 36.95 m, and therefore a rise -to-span ratio of 1/6.77.

Further Reading
Manterola, J., Martínez, A., Navarro, J.A., et al. (2008): Puente de ferrocarril de alta
velocidad sobre el embalse de Contreras. Presented at: IV Congreso ACHE, Valencia.
Manterola, J., Martínez, A., Navarro, J.A., and Martín, B. (2012). Puente arco de
ferrocarril sobre el embalse de Contreras en la línea de alta velocidad
Madrid-Levante. Revista Hormigón y Acero 63 (264).
B.3 Spain 301

Figure B.33 Contreras Bridge (Courtesy of Carlos Fernández Casado CFC & ADIF).

Figure B.34 Contreras Bridge (Courtesy of Carlos Fernández Casado CFC & ADIF).
302 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.6 Viaduct Over River Ulla


B.3.6.1 Data Summary
Owner ADIF
Place Catoira, Spain
HSR Line Atlantic Axis Pontevedra – La Coruña
Designer IDEAM
Main span 240 m
Contractor DRAGADOS-TECSA
Deck Width 14 m
Length 1620 m
Deck type Continuous composite deck
Material Steel
Typology Steel Truss
Start construction 2011
Completion 2015

B.3.6.2 Description
The viaduct over the river Ulla where it flows into the Arosa Estuary. The design
minimises the number of piers on the course of the river and tries to seek the max-
imum transparency and the minimum visual impact possible on the surrounding
landscape. The bridge is a composite variable depth structure with three large cen-
tral spans over the water measuring 225 + 240 + 225 m and 120 m access spans. The
deck has a variable depth with 17.90 m on the piers section and 9.15 m at the cen-
tre of the span. This depth remains constant on the access spans. The four main
piers were reinforced to resist deck rotation and control the level flexion transmit-
ted to the foundations through the frame effect, thus preventing over-sizing. For this
reason, the main piers located at the outer edges of the 225 m spans were designed
with two free-standing partitions driving into the foundations and pier capitals. The
remaining piers are conventional. The deck support over these piers is free spherical
lengthwise bearings. The track has expansion joints coinciding with the expansion
structural joint in the abutments.

Further Reading
Millanes, F., Ortega, M., and Matute, L. (2014). Viaduct over river Ulla: an outstanding
composite (steel and concrete) high-speed railway viaduct. Structural Engineering
International 24.
Millanes, F., Ortega, M., and Estévez, R. (2015). Viaduct over Ulla River in the Atlantic
high-speed railway line: A composite (steel–concrete) truss world record. (ACHE,
ELSERVIER, Hormigón y Acero 66(277).
B.3 Spain 303

Figure B.35 Ulla River Viaduct (Courtesy of ADIF).

Figure B.36 Ulla River Viaduct (Courtesy of ADIF).


304 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.7 Almonte Bridge


B.3.7.1 Data Summary
Owner ADIF
Place Almonte River, Cáceres, Spain
HSR Line Madrid – Extremadura – Portuguese border
Designer Arenas y Asociados
Contractor FCC Fomento de Construcciones y Contratas
Main span 384 m
Deck Width 14.2 m
Length 996 m
Deck type Continuous box girder
Material Concrete
Typology Arch
Start construction 2011
Completion 2016

B.3.7.2 Description
The bridge has a large concrete arch with an upper deck spanning 384 m over the
Alcantara reservoir. This large arch is the main element of a 996 m long viaduct,
consisting of 12 approach spans with 45 m spans, and two additional spans of 36 m
at the ends.
The deck has a prestressed concrete box section with a constant 3.1 m depth. The
viaduct piers have a maximum height of 65.3 m. Both those found at ground and
those supported on the arch have a variable octagonal cross-section, the aerodynam-
ics of which are beneficial for the arch span, given its large span. Almonte’s main
mechanism for taking the longitudinal forces from trains braking is the fixed point
located at the apex of the arch. At the centre of the bridge there is a 42 m long fixed
point connecting the arch and the deck. Horizontal braking loads are transferred
through the fixed point from the deck to the arch and then into the abutments. All
the columns are connected to the deck with bearings, allowing the deck to move
longitudinally with respect to the columns.

Further Reading
Arenas, J.J., Capellán, G., Martínez, J. et al. (2016). Viaduct over River Almonte. Design
and Analysis. In: Presented at: IABSE Symposium: Challenges in Design and
Construction of an Innovative and Sustainable Built Environment, Stockholm,
Sweden.
Capellán, M. (2015). Puente arco de alta velocidad sobre el río Almonte. ROP Revista de
Obras Públicas n∘ 3562 Madrid. Spain.
B.3 Spain 305

Figure B.37 Almonte Bridge (Courtesy of Arenas Asociados & ADIF).

Figure B.38 Almonte Bridge (Courtesy of Arenas Asociados & ADIF).


306 Appendix B Singular Bridges for High-Speed Railway Lines

B.3.8 Alcántara Bridge


B.3.8.1 Data Summary
Owner ADIF
Place Alcántara, Cáceres, Spain
HSR Line Madrid – Extremadura – Portuguese border
Designer Carlos Fernández Casado CFC
Contractor COPISA-COPASA
Main span 324 m
Deck Width 14.2 m
Length 1488 m
Deck type Continuous box girder
Material Concrete
Typology Arch
Start construction 2015
Completion 2019

B.3.8.2 Description
Located on the Madrid-Extremadura line, the viaduct has a total length of 1488 m.
The distribution of the spans of the viaduct is influenced by the Tagus River jump,
which is made by means of an arch of 324 m span, with the deck being divided into
six 54 m spans. The access spans are 60 m, between them are two 57 m transition
spans, one on each side of the springs of the arch. The arch is a concrete hollow
box with a variable edge between 3.50 and 4.00 m and a width between 12.00 m at
the and 6.00 m at the keystone. And the deck is supported by bearings on the piers
and abutments. The keystone of the arch is therefore the point at which the deck is
fixed longitudinally. There are two structural and track expansion joints at the two
abutments.

Further Reading
Manterola, J., Astiz, M.A., and Martínez, A. (1999). Puentes de Ferrocarril de Alta
Velocidad Revista de Obras Públicas n∘ 3386 Madrid.
Manterola, J., Martínez, A., and Martín, B. (2015). Viaducto sobre el río Tajo en el
embalse de Alcántara para ferrocarril de alta velocidad. Revista de Obras públicas
n∘ 3562 Madrid.
B.3 Spain 307

Figure B.39 Alcántara Bridge (Courtesy of Carlos Fernández Casado CFC & ADIF).

Figure B.40 Alcántara Bridge (Courtesy of Carlos Fernández Casado CFC & ADIF).
B.4 Japan 309

B.4 Japan
The first high-speed lines were built in Japan. The population density throughout
Japan has meant that most of the lines have been built on viaducts. The typologies of
bridges for the Shinkansen are varied, although the most singular are the extradosed
bridges, a typology that originated in Japan itself.
310 Appendix B Singular Bridges for High-Speed Railway Lines

B.4.1 Yashiro Bridge


B.4.1.1 Data Summary
Owner East Japan Railway Company
Place Chikuma, Nagano, Japan
HSR Line Hokuriku Shinkansen
Designer Japan Bridge & Structure Institute, Inc.
Contractors JV of the Zenitaka Corp.; Daiho Corp.; Moriya Corp.;
Sumitomo Mitsui Construction Co. Ltd.
Main span 2 × 105.0 m
Deck Width 12.8 m
Length 696.0 m
Deck type Box girder
Material Concrete
Typology Extradosed
Start construction 1994
Completion 1996

B.4.1.2 Description
The bridge consists of several sections, including a continuous extradosed section
with spans of 65 + 105 + 105 + 105 + 65 m. The deck is a prestressed concrete box
with a depth of 2.50 m and is connected monolithically to the shaft of the piers,
which have a height of 12 m above the deck. This bridge is the first application of
the extradosed bridge typology on a high-speed railway line.

Further Reading
Yuyama, K.Y. and Watanabe, J. (1998). Innovative new type of cable-stayed bridge -
Yashiro bridges of Hokuriku Shinkansen line it. In: Prestressed Concrete in Japan.
1998. XIII FIP Congress National Report, Amsterdam, Holland.
https://structurae.net/en/structures/yashiro-bridge.
B.4 Japan 311

Figure B.41 Yashiro Bridge (Courtesy of the Zenitaka Corporation).

Figure B.42 Yashiro Bridge (Courtesy of the Zenitaka Corporation).


312 Appendix B Singular Bridges for High-Speed Railway Lines

B.4.2 Kumagawa Bridge


B.4.2.1 Data Summary
Owner Kyushu Railway Company
Place Yatsushirogun, Kumamoto Prefecture, Japan
HSR Line Kyushu Shinkansen
Designer Japan Bridge & Structure Institute, Inc.
Contractor Joint Venture of The Zenitaka Corp., Daisue Corp.
Main span 35.0 + 62.0 + 75.0 + 82.0 + 53.0 m
Deck Width 11.3 m
Length 307.0 m
Deck type Box girder
Material Concrete
Typology Frame
Start construction 2000
Completion 2002

B.4.2.2 Description
It is a continuous monolithic bridge with foundations by means of caissons
of 11.00–12.5 m in diameter and 10–16 m in vertical dimension. The deck is a
pre-stressed concrete box with a variable depth that is connected monolithically to
the piers. The track is ballastless.
B.4 Japan 313

Figure B.43 Kumagawa Bridge (Courtesy of the Zenitaka Corporation).

Figure B.44 Kumagawa Bridge, view from below (Courtesy of the Zenitaka Corporation).
314 Appendix B Singular Bridges for High-Speed Railway Lines

B.4.3 Sannai-Maruyama Bridge


B.4.3.1 Data Summary
Owner East Japan Railway Company
Place Aomori, Japan
HSR Line Tohoku Shinkansen
Structural Engineer Chiyoda Engineering Consultants
Contractor JV of the Zenitaka Corp.; Asanuma Corp.; Shida Corp.
Main span 74.18 + 150 + 150 + 74.18 m
Deck Width 13.85 m
Length 450.0 m
Deck type Box girder
Material Concrete
Typology Extradosed
Start construction 2006
Completion 2008

B.4.3.2 Description
The Bridge on the Tohoku Shinkansen extension was constructed to cross a broad
highway, river, and reservoir near the famous Sannai-Maruyama archaeological
site (Figure 2). It consists of two 75 m spans and two 150 m spans of extradosed
bridges. The bridge with 4 continuous spans has a pneumatic caisson foundation.
The total length is 450 m with spans: 74.18 + 150 + 150 + 74.18 m. The height of
Pylon is 17.5 m, and the material of the stay cable is Epoxy-coated strand. The
bridge features saddles in the pylon in which the strands run through a curved steel
pipe The installation of sliding elastic bearings on four piers except the centre pier,
which is the stationary point of this bridge, reduces the transfer of displacements
into the piers and girder due to expansion and contraction of the girder caused by
seasonal thermal fluctuation.

Further Reading
DYWIDAG Systems International (ed.) (2009–2010). Moderne Technik und Tradition:
Extradosed-Brücke in Japan. In: DSI Info, vol. 17.
Tamai, S. and Shimizu, K. (2011). The long spanned bridge for deflection-restricted
high-speed rail Sannai-Maruyama Bridge. In: Presented in 9th World Congress on
Railway Research Lille France.
https://structurae.net/en/structures/sannai-maruyama-bridge.
B.4 Japan 315

Figure B.45 Sannai-Maruyama (Courtesy of the Zenitaka Corporation).

Figure B.46 Sannai-Maruyama, aerial view (Courtesy of the Zenitaka Corporation).


B.5 China 317

B.5 China
The long viaducts currently being built in China are isostatic sections that are pre-
fabricated in their entirety and assembled directly from previously built sections.
In addition, the orography of mountainous areas and the width of riverbeds have
led to the construction of long-span bridges: arches, cable-stayed bridges and even
the first suspension bridge are undoubtedly a sign of China’s determination to build
a high-speed network that does not shy away from any technical challenge.
318 Appendix B Singular Bridges for High-Speed Railway Lines

B.5.1 Tianxingzhou Yangtze River Bridge


B.5.1.1 Data Summary
Owner China Railway
Place Hubei Province, China
HSR Line Beijing – Guangzhou
Main span 504 m
Deck Width 27 m
Length 4657.1 m
Deck type Steel truss girder
Material Steel
Typology Cable stayed bridge
Start construction 2003
Completion 2009

B.5.1.2 Description
The bridge deck has two levels: the upper one for a motorway with 3 lanes in each
direction and the lower one for 4 high-speed tracks. The deck consists of three truss
planes, braced together, which are braced by three cable planes to the two main
pylons. The main lateral offset spans have a short span adjacent to them to reduce
rotations at the expansion joints of the structure. The railway tracks are ballasted.

Further Reading
Qin, S. and Gao, Z. (2017). Developments and prospects of long-span high-speed
railway bridge technologies in China. Engineering 3 (6): https://doi.org/10.1016/j.eng
.2017.11.001.
B.5 China 319

Figure B.47 Tianxingzhou Yangtze River Bridge (Courtesy of China Railway).

Figure B.48 Tianxingzhou Yangtze River Bridge (Courtesy of China Railway).


320 Appendix B Singular Bridges for High-Speed Railway Lines

B.5.2 Nanjing Dashengguan Yangtze River Bridge


B.5.2.1 Data Summary
Owner China Railway
Place Jiangsu Province, China
HSR Line Beijing – Shanghai
Designer China Railway Major Bridge Reconnaissance &
Design Institute Co.
Main span 336 m
Deck Width 41.6 m
Length 1615 m
Deck type Steel truss girder
Material Steel
Typology Arch + Beam
Start construction 2006
Completion 2011

B.5.2.2 Description
The main bridge has a total length of 1.615 m and is the longest truss span built
to date. Transversally, the bridge is made up of three trussed beams, braced against
each other, which accommodate the six railway tracks that run along the bridge. The
hangers are rigid steel tubes with a hollow rectangular cross-section of 1.00 × 1.40 m.
The edges of the rectangle formed by the rigid hangers are chamfered to eliminate
possible aeroelastic vibrations. The tracks are ballasted throughout the structure.

Further Reading
Gao, Z., Yi, L., and Xiao, H. (2009). Dashengguan Bridge – the longest span arch bridge
for high-speed railway. In: Presented at: IABSE Workshop: Recent Major Bridges,
Shanghai, China.
Gao, Z., Yi, L., and Xiao, H. (2010). Dashengguan Bridge—the largest span steel arch
bridge for high-speed railway. Structural Engineering International 20 (3).
https://structurae.net/en/structures/dashengguan-bridge.
B.5 China 321

Figure B.49 Nanjing Dashengguan Yangtze River Bridge (Courtesy of China Railway).

Figure B.50 Nanjing Dashengguan Yangtze River Bridge (Courtesy of China Railway).
322 Appendix B Singular Bridges for High-Speed Railway Lines

B.5.3 Tongling Yangtze River Bridge


B.5.3.1 Data Summary
Owner China Railway
Place Anhui Province, China
HSR Line Hefei – Fuzhou
Main span 630 m
Deck Width 34.2 m
Length 1290 m
Deck type Steel truss girder
Material Steel
Typology Cable stayed bridge
Start construction 2010
Completion 2015

B.5.3.2 Description
The bridge deck has two levels: the upper for a motorway with 3 lanes in each direc-
tion and the lower for 4 high-speed tracks. The deck consists of three truss planes,
braced together, which are braced by three cable planes to the two main pylons. The
main lateral offset spans have a short span adjacent to them to reduce rotations at
the expansion joints of the structure. The railway tracks are ballasted.

Further Reading
Qin, S. and Gao, Z. (2017). Developments and prospects of long-span high-speed
railway bridge technologies in China. Engineering 3 (6): https://doi.org/10.1016/j.eng
.2017.11.001.
https://structurae.net/en/structures/tongling-road-rail-bridge.
B.5 China 323

Figure B.51 Tongling Yangtze River Bridge (Courtesy of China Railway, Baidu).

Figure B.52 Tongling Yangtze River Bridge (Courtesy of China Railway).


324 Appendix B Singular Bridges for High-Speed Railway Lines

B.5.4 Beipanjiang Bridge


B.5.4.1 Data Summary
Owner China Railway
Place Guizhou Province, China
HSR Line Shanghai – Kunming
Main span 445 m
Deck Width 13.4 m
Length 721 m
Deck type Box girder
Material Concrete
Typology Arch + Beam
Start construction 2010
Completion 2016

B.5.4.2 Description
The main structure is a concrete arch bridge with a span of 445 m and a rise of 100 m,
giving a rise-to-span ratio of 1/4.45. The arch cross-section is a box containing three
cells, with a constant depth of 9.0 m, but varying in width and wall thickness to
improve the transverse stability for the structure. This is important because the
width of the bridge deck is only 13.4 m resulting in a small width-to-span ratio in
this super-long span arch. In each adjacent pier, the two 65-m girders are rigidly
connected to the pier and form a T-shaped rigid frame structure. The overall super-
structure is in the form of a (2 × 65 m + 8 × 42 m + 2 × 65 m) prestressed concrete
rigid-frame/ continuous girder with a total length of 599.6 m. Such a continuous
structure would effectively improve the vertical and transverse stiffness and help
assure effective operation of the high-speed train. The main arch was constructed
by the embedded Concrete-Filled Steel Tube (CFST) scaffolding method. The
40-segment, 445-m-span steel tube truss arch was first erected, then hoisted by
cable crane, and erected by the cable-stayed cantilever method. Concrete grade of
C80 was injected into the steel tubes after the closure of the steel tube truss arch to
make the truss arch a CFST structure, resulting in significant improvement in the
stiffness and strength of the trussed arch.

Further Reading
Chen, B., Su, J.Z., Lin, S. et al. (2017). Development and application of concrete arch
bridges in China. Journal of Asian Concrete Federation 3 (1): 12–19.
Chen, B., Liu, J., and Tabatabai, H. (2019). Recent research and application of arch
bridges in China. In: 9th International Conference on Arch Bridges. Porto, Portugal.
B.5 China 325

Figure B.53 Beipanjiang Bridge (Courtesy of China Railway).

Figure B.54 Beipanjiang Bridge (Courtesy of China Railway).


326 Appendix B Singular Bridges for High-Speed Railway Lines

B.5.5 Yachihe Bridge


B.5.5.1 Data Summary
Owner China Railway
Place Sichuan Province, China
HSR Line Chengdu – Guiyang
Main span 436 m
Deck Width 22 m
Length 971 m
Deck type Box girder
Material Concrete
Typology Arch + Beam
Start construction 2014
Completion 2018

B.5.5.2 Description
The Yachihe Railway Bridge is nearly 300 m over the reservoir and has ac main arch
span of 436 m. The twin track structure is composed of a steel–concrete composite
truss. The approach spans on either side of the main arch are composed of a T-beam
of 2 × 61.75 m.

Further Reading
Chen, B., Liu, J., and Tabatabai, H. (2019). Recent research and application of arch
bridges in China. In: 9th International Conference on Arch Bridges. Porto, Portugal.
B.5 China 327

Figure B.55 Yachihe Bridge (Courtesy of China Railway).

Figure B.56 Yachihe Bridge (Courtesy of China Railway).


328 Appendix B Singular Bridges for High-Speed Railway Lines

B.5.6 Wufengshan Yangtze River Bridge


B.5.6.1 Data Summary
Owner China Railway
Place Suzhong Province, China
HSR Line Lianyungang–Zhenjiangand and
Jiangyi Expressway
Main span 1092 m
Deck Width 41 m
Length 1428 m
Deck type Steel truss girder
Material Steel
Typology Suspension Bridge
Start construction 2015
Completion 2020

B.5.6.2 Description
The Wufengshan Yangtze River Bridge is designed for both road and rail use, with
a total length of 6.409 km and a main bridge with a single span of 1092 m. A 4-lane
railway plus a two-way 8-lane expressway. It is the world’s first road-high-speed rail
suspension bridge. The distribution of the spans is: 84 + 84 + 1092 + 84 + 84 m.

Further Reading
Qin, S. and Gao, Z. (2017). Developments and prospects of long-span high-speed
railway bridge technologies in China. Engineering 3 (6). https://doi.org/10.1016/j.eng
.2017.11.001.
B.5 China 329

Figure B.57 Wufengshan Yangtze River Bridge (Courtesy of China Railway).

Figure B.58 Wufengshan Yangtze River Bridge (Courtesy of China Railway).


331

Index

a b
abutments, of bridge 90, 108, 112, 113 ballasted track 186, 187
access on 241–242 definitions, functions, and qualities
continuous deck 32
with dampers located on 223 evolution of 36–37
with single fixed point at one longitudinal design 65–66
203–211 optimised 41–43
with expansion joint partial ballast removal 119, 120
only in structure 90 ballastless track 43–45, 66
in structure and track 90–92 bearings, for bridges
fixed loads arrangement continuous decks 97, 99
forces acting on 138–140 friction 124
loads transmitted by deck 137–138 layout 112
transversal cross section of 137 POT bearings 95, 98, 112
fixed point 91–95 replacement 242, 243
accidental situations 118 simply supported bridges 96–98, 224,
aerodynamic actions, from passing trains 226
123 spherical bearing 96, 98
Alcántara Bridge 20, 306, 307 track-structure interaction 189–190
alert limit (AL) 47 Beipanjiang Bridge 18–20, 324, 325
alignment, track geometric quality 47 bi-block sleepers 37, 38
Almonte Bridge 20, 304, 305 braking force 121–122
antimetric vibration modes 169–171 bridge substructure elements, impact on
anti-noise panel, on bridge 14 126
AP-7 Viaduct Almeria, Spain 82 British high-speed bridges 12
arch bridges Built Bridges 66
tied 85, 86
upper deck bridges 83–85 c
Arroyo de las Piedras Viaduct 223 cable-stayed bridges 13, 86–88
asymmetric switch rail profile 197, 198 cable-supported bridges
Avignon viaducts 8, 284, 285 cable-stayed bridges 86–88

High-Speed Railway Bridges: Conceptual Design Guide, First Edition.


José Romo, Alejandro Pérez-Caldentey, and Manuel Cuadrado.
© 2024 Ernst & Sohn GmbH. Published 2024 by Ernst & Sohn GmbH.
332 Index

cable-supported bridges (contd.) hybrid bridges 89


extradosed bridges 85–86 suspension bridges 88
hybrid bridges 89 deck pre-dimensioning 109–111
suspension bridges 88 design situation 66–67
central fixation 71–72 high-level viaducts 71–72
central piers long span structures 72
continuous deck long viaducts 67–71
with fixed point at one 211 low profile 68–71
with several fixed points at 221, 222 short crossing at low level 67
for fixed point, in Spanish HS railway fixed point, of bridge 108
220 information analysis 107–108
centrifugal forces 122, 132, 134–135 infrastructure pre-design
characteristic equation, dynamic analysis abutments 112, 113
174–177 fixed point and bearings 112
China bridges piers 113–114
Beipanjiang Bridge 18–20, 324, 325 longitudinal design strategies
description of 317 ballasted track 65–66
high-speed railway bridges 10, 11 ballastless track 66
Hutong Yangtze River Bridge 21–22 fixed points, design of 66
isostatic deck 17 short and long viaduct 64
isostatic multi span bridge 10, 11 simply supported decks 64
isostatic solutions 13 ultra-long and continuous viaducts
Nanjing Dashengguan Yangtze River 64, 65
Bridge 320, 321 methodology 107
Tianxingzhou Yangtze River Bridge overview of 61
318, 319 plan and elevation 106
Tongling Yangtze River Bridge 322, preliminary geotechnical data 107
323 railway platform 106
Wufengshan Yangtze River Bridge 88, seismic design (see seismic design)
328, 329 span distribution 109
Yachihe Bridge 326, 327 straight deck solutions 72
Colne Valley Viaduct, England 12, 15 concrete box hollow sections 78–80
columns, track-structure interaction 190 continuous slab concrete decks
Complete Quadratic Combination (CQC) 73–75
155 precast beam decks 74–77
composite bridges, with double ‘I’ steel simply supported solutions 72–73
beam 80, 81 steel beam decks 80–81
concentrated mass matrix 171–172 steel semi-through decks 81–82
conceptual design structural and functional requirements
arch bridges braking and traction forces 62–63
tied 85, 86 expansion joints 62, 63
upper deck bridges 83–85 longitudinal fixation 62
cable-supported bridges track to bridge deck interaction 63
cable-stayed bridges 86–88 vertical accelerations control 62
extradosed bridges 85–86 vertical stiffness 62
Index 333

structural elements drainage of 247


abutments 90–95 expansion joints 243–246
bearings 95–99 impact against structural elements of
piers 95–97 125–126
topography 106 isostatic 7, 17
truss bridges 82, 83 loaded state 134
water flood level 107 longitudinal displacement of 54–55
concrete box hollow sections 78–80 longitudinal schemes (see longitudinal
Concrete-Filled Steel Tube (CFST) track-bridge interaction)
scaffolding method 324 precast beam 74–77
concrete slab deck type, solid pier for 95, pre-dimensioning 109–111
96 simply supported (see simply supported
continuous bridges 17–19 deck)
continuous decks steel beam 80–81
with dampers located on 223 steel semi-through 81–82
longitudinal behaviour, seismic design track-structure interaction 189
100–101 track to deck interaction 63
seismic design 220–223 transverse deformation and vibration of
with single fixed point at one 203–211 53–54
Spanish bridge with 8, 9 twist 51
track-structure interaction 220–223 unloaded state 133
continuous frame 69 vertical acceleration of 48–50
continuous slab concrete decks 73–75 vertical deformation of 51–53
continuous viaduct 71 wind force in 133, 134
continuous welded rail (CWR) 35–36, deck twist 51
185, 186 derailment 125
Contreras Bridge 300, 301 design regulations, for bridges
conventional vs. high-speed railway accidental load situations 117
bridges 4 cross section 129
creep and shrinkage 192–193 deformation and vibrations,
crossbar control lever mechanism 199 verifications 128–129
cross-level, track geometric quality 47 dynamics effects
CWR see continuous welded rail (CWR) consideration of 125
principal factors 124–125
d railway vehicle derailment 125–126
damping devices plus bearings 104–105 static forces 124
damping index 180, 181 elevation 129
dead loads 110, 111, 118–119 fixed abutment, forces and preliminary
deck (bridge) 72 design of 130
access on 240 fixed abutment loads
concrete box hollow sections 78–80 forces acting on 138–140
continuous (see continuous decks) loads transmitted by deck 137–138
continuous slab concrete decks 73–75 transversal cross section of 137
cross section of 110, 111 horizontal forces
double-level truss 87 centrifugal force 132
334 Index

design regulations, for bridges (contd.) dynamic amplification


traction and braking 131–132 of deck deformation and acceleration
permanent loads 49
dead loads 118–119 and resonance 2–3
partial ballast removal 119, 120 dynamic analysis see also Dynamic Load
self-weight 118 Factor (DLF)
pre-dimensioning 130 antimetric vibration modes 169–171
centrifugal forces 134–135 characteristic equation 174–177
loads per bearings 136 concentrated mass matrix 171–172
pier heads, reaction in 135 damping effects 164–165
transversal wind bearings reactions damping index 180, 181
136 degrees of freedom 169
vertical forces 133–134 diagonal matrix 172–173
wind at unloaded state 135 eigenfrequency and eigenmode
wind with live load 135 180
project situations 117–118 eigenvectors matrix 178–179
traffic loads interoperability 167–168
application of 126–128 isolated load 158–161
for fatigue 128 maximum deflection 182–183
variable loads modal superposition method 153
aerodynamic actions, from passing eigenforms, formulation based on
trains 123 155–158
bearings’ friction 124 matrix formulation, for finite
centrifugal forces 122 element analysis 153–155
nosing forces 123 resonance 143–144
thermal actions 123 stiffness matrix 170
traction and braking forces symmetric vibration modes
121–122 169–171
vertical live loads 119–121 train load 162–164
vertical loads vehicle-structure interaction model
dynamic factor 131 165–167
eccentricity of 131 dynamic load factor (DLF)
live loads 130–131 damping effects 165
permanent loads 130 defined 143
wind speed 132–134 dynamic analysis 147–152
diagonal matrix 172–173 envelope dynamic factor 144–145,
discontinuous deck, seismic design 168–169
223–226 for real trains 145–147
displacements, verification of single degree-of-freedom system
horizontal displacement 195–196, 249–257
231–234 train velocity, function of 164
vertical displacement 196, 234–235 dynamics effects, of bridge
DLF see Dynamic Load Factor (DLF) consideration of 125
double-level truss deck 87 derailment 125
drainage maintenance 246–247 principal factors 124–125
Index 335

railway vehicle impacts finite element analysis, matrix


on bridge substructure elements formulation for 153–155
126 first high-speed bridges 6
structural elements, of deck French bridges 8
125–126 German bridges 6–7
static forces 124 Spanish bridges 8–9
‘Fish belly’ rail 40
e fixed abutments
earthquake intensity levels 225 with brackets 92, 93
Eigenforms, formulation based on with friction slab 94, 95
155–158 longitudinal configuration of 91, 92
elastic direct fastener 39 with POT bearings 91–94
elastic indirect fastener 39, 40 with prestress 92, 94
EN 1337-1 124 fixed transversal support 101–102
EN-1990 127, 128 force per length vs. longitudinal
EN 1991-2 50–52, 54, 121–123, 125, 127, displacement law 189, 190, 208,
130, 145, 148, 167, 189, 191, 194 226
foundations, track-structure interaction
EN 1990 Annex A2 52, 53, 56
190–191
England, Colne Valley Viaduct in 15
frame sleepers 38, 39
envelope dynamic factor 144–145,
French bridges
168–169
Avignon viaducts 284, 285
ERRI committees 55
description of 281
ERRI D190 committee 57
first-generation 8
Europe
Garde-Adhémar Viaduct 282, 283
aesthetics in 22
Mornas Viaduct 286, 287
continuous bridges 13
Savoureuse Viaduct 9, 17, 288, 289
European railways systems,
Fresno Viaduct, in California 12
interoperability of 167
exceptional geometries, track-structure g
interaction 226–228 Gänsebachtal Viaduct 5, 6, 9, 10, 72,
expansion joints 274, 275
drain trough 247 Garde-Adhémar Viaduct 282, 283
example of 244, 246 Gemünden Bridge 7, 18, 264, 265
neoprene expansion joint 244, 246 German bridges
rotation at 62 description of 263
steel plates joint Filstal Bridge 278, 279
in fixed abutment 243, 244 first-generation 6–7
in movable abutment 244, 245 Gänsebachtal Viaduct 5, 6, 9, 10, 18,
track 63 19, 72, 274, 275
extradosed bridges 85–86 Gemünden Bridge 18, 264, 265
Hämerten Bridge 276, 277
f Nantenbach Bridge 21, 270, 271
fasteners 33, 39–40 Pfieffetal Bridge 18, 268, 269
Filstal Bridge 278, 279 semi-integral bridges 13
336 Index

German bridges (contd.) Sannai-Maruyama Bridge 314, 315


Unstruttal Bridge 9, 10, 18, 19, 272, Yashiro Bridge 310, 311
273
Veitshöchheim Bridge 7, 266, 267 k
German Requirements Catalogue 200 Kumagawa Bridge 312, 313
grooved rail 41
l
h La Grenette Viaduct 8
Hämerten Bridge 276, 277 landscape
heavy load models 120 bridges in 15–22
high-level viaducts 71–72 traveller’s 13–15
High Speed Load Model (HSLM-A) 50, Lisbon, Tagus River Crossing in 22, 87,
148, 167 88
geometry of 148, 149 Llinars Del Vallès Viaduct 294, 295
maximum acceleration 150, 151 longitudinal displacement, of deck
maximum displacement 150, 151 54–55
parameters of 148, 149 longitudinal level
hollow pier, with variable dimensions defined 47
95, 97 limit values for 47, 48
HSLM-A see High Speed Load Model longitudinal track-bridge interaction
(HSLM-A) 203
HSLM-B trains 167–168 continuous deck
Hutong Yangtze River Bridge, China with fixed point, one central piers
21–22 211
hybrid bridges 89 with single fixed point, one
hydraulic dampers, deck and abutment abutments 203–211
100 deck divided into continuous stretches
hydraulic viscodamper 104, 105 216–218
fixed points at two abutments and
i structural joint in middle
immediate action limit (IAL) 47 212–216
International Union of Railways simply supported spans 211–213
publication UIC 776-1 145 long-span bridges 11, 13, 23
interoperability 167–168
intervention limit (IL) 47 m
inverted frame system 69–70 maintenance, conceptual design for
isolated load 158–162 accesses 240
isostatic bridges 17 abutments 241–242
isostatic multi span bridge, China decks 240
10, 11 piers 240–241
bearings 242–243
j designers work 239
Japan bridges drainage 246–247
description of 309 expansion joints 243–246
Kumagawa Bridge 312, 313 optimal structures 239
Index 337

Maria Pia Bridge, Portuguese 23–24 elastic re-centring element 105


mean wind speed 132 foundations of 113–114
Millipede bridges 69 transversal damping system in 102
modal superposition method 153 transversal stoppers at 102
eigenforms, formulation based on Portuguese, Maria Pia Bridge 23–24
155–158 pot-bearing behaviour model 223, 224
matrix formulation, for finite element precast beam decks 74–77
analysis 153–155 pre-fabricated solution 76
monobloc sleepers 37, 38 Prud’homme formula 45
monotonous bridge 69
Mornas Viaduct 286, 287 r
movable scaffolding systems (MSS) 79, rail expansion joints (REJs)
80 bridge needed 204–206
multicomponent action, characteristic bridge not needed 206–211
values of 127–128 deck divided into continuous stretches
multiple degrees of freedom (MDOF) 216–218
153 design of
asymmetric switch rail profile 197,
n 198
Nanjing Dashengguan Yangtze River crossbar control lever mechanism
Bridge 320, 321 199
Nantenbach Bridge 21, 270, 271 functional principle of 197
Nationally Determined Parameters (NDP) gaps 199, 200
189 maximum displacement 197
neoprene expansion joint 244, 246 picture of 198
nosing forces 123 uplift forces 200
installation 201–204
o rail stresses with 215, 216
Osera Bridge 292, 293 regulation 201, 202
rails
p track component 33–34
passenger comfort 3, 55–57 ‘fish belly’ 40
peak wind speed 132 grooved 40, 42
pendulum devices 103, 104 Vignoles type 40, 41
permanent loads 130 track-structure interaction 189
dead loads 118–119 rail stresses
partial ballast removal 119, 120 braking load 210
self-weight 118 computed for bridge 207, 213
persistent situations 117 envelope of 211, 228
Pfieffetal Bridge 7, 18, 268, 269 for left viaduct 218
pier heads with REJ 215, 216
maximum axial reaction in 136 for right viaduct 218
reactions in 135 seasonal variation of temperature 214
piers 95–97 vertical traffic load 210
access on 240–241 rail traffic security 3
338 Index

Real Trains 50 Shock Transmitter Units (STUs) 222,


REJs see rail expansion joints (REJs) 225
Riudellots Viaduct, Spain 67 short-span bridges 22
simply supported deck 72–73, 96–98
s bearings for 224, 226
Salto Del Carnero Railway Bridge 296, longitudinal behaviour, seismic design
297 100
Sannai-Maruyama Bridge 314, 315 without longitudinal continuity
Sar Viaduct, Spain 4, 5 211–213
Savoureuse Viaduct 9, 17, 288, 289 single degree-of-freedom (SDOF) system
SDOF system 150, 152, 154 see also damped system, subjected to triangular
single degree-of-freedom (SDOF) load 258
system auxiliary expressions 259–261
seismic design time interval 261–262
longitudinal behaviour Dynamic Load Factor 249–257
continuous decks 100–101 resonance 257–258
simply supported spans 100 singular bridges 18–22
seismic behaviour and deck articulation sleepers 32–33, 37–39
99–100 SLS see Serviceability Limit State (SLS)
strategies 99 span by span isostatic solution (China)
track-structure interaction 218–220 5
continuous deck 220–223 Spanish bridges
discontinuous deck 223–226 Alcántara Bridge 20, 306, 307
transversal behaviour Almonte Bridge 20, 304, 305
damping devices plus bearings AP-7 Viaduct Almeria 82
104–105 Contreras Bridge 300, 301
fixed transversal support 101–102 description of 291
relative vertical rotation axis 101 first-generation 8–9
transversal damping systems Llinars Del Vallès Viaduct
102–104 294, 295
seismic situations 118 Osera Bridge 292, 293
Serviceability Limit State (SLS) 117 recent bridges 10
for bridge 31 Riudellots Viaduct 67
deck Salto Del Carnero Railway Bridge 296,
longitudinal displacement of 297
54–55 Ulla Bridge 21, 24
transverse deformation and vibration Viaduct Over AP7 Riudellots de la Selva
of 53–54 298, 299
twist 51 Viaduct Over River Ulla 302, 303
vertical acceleration of 48–50 Spanish HS railway, central pier for fixed
vertical deformation of 51–53 point in 220
levels of comfort 55–57 Square Root of the Sum of the Squares
track geometry quality 46–48 (SRSS) 154, 161
track-vehicle dynamic interaction steel beam decks 80–81
44–46 steel damping bearings 103, 104
Index 339

steel plates joint track gauge, defined 47


in fixed abutment 243, 244 track geometry quality 46–48
in movable abutment 244, 245 track-infrastructure interaction 36
steel semi-through decks 81–82 track joints 187, 188, 191, 196
steel sleepers 37, 38 track-structure interaction 4, 123
steel truss bridges 82, 83 analysis, model for 188
stiffness matrix 170 bearings 189–190
STUs see Shock Transmitter Units (STUs) columns 190
suspension bridges 88 force per length vs. longitudinal
switches/crossings 34–35 displacement 189, 190, 208, 226
symmetric vibration modes 169–171 foundations 190–191
horizontal levels 188
t rails and deck 189
Tagus River Crossing, in Lisbon 22, 87, creep and shrinkage 192–193
88 description of 185
temperature variations 191 exceptional geometries 226–228
thermal actions 123 load combinations for 194
Tianxingzhou Yangtze River Bridge 318, longitudinal schemes 203
319 continuous deck with fixed point,
tied arch bridges 85, 86 one central piers 211
timber sleeper 37 continuous deck with single fixed
Tongling Yangtze River Bridge 322, 323 point, one abutments 203–211
track, for high-speed bridges deck divided into continuous
ballasted 65–66 stretches 216–218
ballastless 43–45, 66 fixed points at two abutments
components 212–216
ballast 32, 36–37 simply supported spans 211–212
fasteners 33, 39–40 problem statement
rails 33–34, 40–41 ballasted track 186, 187
sleepers 32–33, 37–39 continuous welded rail 185, 186
switches and crossings 34–35 relative displacements 187
continuous welded rail 35–36 relative movements 186
conventional track superstructure 31 structural configuration 187–188
dead loads, road vs. railway bridges vertical traffic loads 187–188
29, 30 rail expansion joints 197–203
description of 29 seismic design 218–220
expansion joints 63 continuous deck 220–223
operational conditions 31 discontinuous deck 223–226
optimised ballasted track 41–43 step-by-step computations 193–194,
SLS (see Serviceability Limit State 209, 210
(SLS)) temperature variations 191
track-infrastructure interaction 36 traction and braking forces 191–192
traffic loads, road vs. railway bridges verifications
29–31 horizontal displacement 195–196,
ULS, track and vehicle 31 231–234
340 Index

track-structure interaction (contd.) Unstruttal Bridge 9, 10, 18, 19, 272, 273
stresses 194–195, 229–234 upper deck arch bridges 83–85
track joints 196 USA high-speed rail bridges 12
vertical displacement 196, 234–235 U-shaped beams 75
vertical loads 192
track-vehicle dynamic interaction 44–46 v
traction and braking forces 121–122 Veitshöchheim Bridge 7, 266, 267
horizontal forces 131–132 verifications, of track-structure
track-structure interaction 191–192, interaction
195 horizontal displacement 195–196,
traffic loads 231–234
application of 126–128 stresses 194–195, 229–234
for fatigue 128 track joints 196
train load 162–164 vertical displacement 196, 234–235
transitional situations 118 vertical acceleration
transversal behaviour, seismic design control of 62
damping devices plus bearings of deck 48–50
104–105 vertical deformation, of deck 51–53
fixed transversal support 101–102 vertical loads 133–134
relative vertical rotation axis 101 dynamic factor 131
transversal damping systems eccentricity 131
102–104 live loads 119–121, 130–131
transversal wind bearings reactions 136 permanent loads 130
transverse deformation and vibration, of POT bearings 136
deck 53–54 track-structure interaction 192
transverse load imbalances, between rails vertical track stiffness 46
121 vertical traffic loads 186–187
traveller’s landscape 13–15 Viaduct Over AP7 Riudellots de la Selva
truss bridges 82, 83 298, 299
21st century, structural engineering of Viaduct Over River Ulla 302, 303
24 vibration modes 161, 162
Vignoles rail 40, 41
u
U cross sections 67–69 w
UIC leaflet 774-3 54 wind speed 132–134
UIC leaflet 776-3 52, 53, 200 Wufengshan Yangtze River Bridge 88,
Ulla Bridge, Spain 21, 24 328, 329
tubular-framed 89
Ultimate Limit State (ULS) 31, 117 y
Universal dynamic train A 167 Yachihe Bridge 326, 327
Universal dynamic train B 167–168 Yashiro Bridge 310, 311
WILEY END USER LICENSE AGREEMENT
Go to www.wiley.com/go/eula to access Wiley’s ebook EULA.

You might also like