07.07.2017 Views

SEI-november-2010-journal

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>SEI</strong> Volume 20 | Number 4 | November <strong>2010</strong><br />

STRUCTURAL ENGINEERING<br />

INTERNATIONAL<br />

International Association for Bridge and Structural Engineering (IABSE)<br />

Fibre Reinforced<br />

Polymer Composites


Bridge Information Modeling –<br />

It’s everything to do with the bridge.<br />

Bridge Information Modeling is a compelling new methodology for project delivery that dramatically improves bridge quality and<br />

reduces risk. Using the Bentley bridge solution for BrIM, engineers develop and employ an unprecedented depth of information<br />

about the bridge as they streamline the entire bridge development process, design through construction engineering.<br />

Organizations achieve the highest level of real-time team collaboration using the Bentley solution. They get the right information<br />

in the right hands and ultimately empower project stakeholders to collectively deliver the best-quality, most constructible bridges<br />

possible. They also deliver an invaluable bridge information asset to the owner for lifetime operation of the bridge.<br />

With our leading bridge design and engineering systems, RM Bridge and LEAP Bridge, projects evolve in a synthesis of modeling,<br />

analysis, and design. Engineers perform an array of intensive analyses of any condition, from seismic impact to wind dynamics and<br />

even rolling stock analysis – for every stage of construction.<br />

Join the world’s top engineering consultancies. Choose the Bentley bridge solution for BrIM – and the satisfaction of delivering safe,<br />

sustainable bridges on time and on budget.<br />

Only Bentley can take you there. www.bentley.com/<strong>SEI</strong><br />

© <strong>2010</strong> Bentley Systems, Incorporated. Bentley, the “B” Bentley logo, and LEAP are either registered or unregistered trademarks or service marks of Bentley Systems, Incorporated or one of its direct or indirect whollyowned<br />

subsidiaries. Other brands and product names are trademarks of their respective owners.


idges - linking people ®<br />

Bandra Worli Sea Link, India<br />

Equipped with 254 DISKTRON disc bearings (Vmax =<br />

14,000 kN) and 26 mageba TENSA ® MODULAR expansion<br />

joints of types LR2 to LR6 (max. movement: 480mm).<br />

Canada<br />

China<br />

Croatia<br />

Denmark<br />

Finland<br />

France<br />

Germany<br />

Hong Kong<br />

Poland<br />

Portugal<br />

Slovenia<br />

South Korea<br />

Switzerland<br />

USA<br />

structural bearings seismic devices<br />

expansion joints monitoring & services<br />

Switzerland www.mageba.ch


Sika’s Advanced Technologies<br />

for Bridge Engineering<br />

Sika ® ViscoCrete ®<br />

High performance concretes using<br />

Sika ® ViscoCrete ® Admixture<br />

Technology<br />

• Self compacting<br />

• Fast /Controlled setting<br />

• Watertight and corrosion resistant<br />

SikaGrout ®<br />

Long lasting cable ducts using<br />

SikaGrout ® PT-Technology<br />

• Factory quality controlled<br />

• Waterproof<br />

• Corrosion resistant<br />

Sika ® CarboDur ®<br />

Structural strengthening using<br />

Sika ® CarboDur ® Technology<br />

• CFRP plate systems<br />

• Fabric systems<br />

• Patented prestressing systems<br />

• CarboShear ® system<br />

Sika Services AG<br />

Business Unit Constractors<br />

CH-8330 Pfäffikon<br />

Switzerland<br />

Phone +41 58 436 23 79<br />

Fax: +41 58 436 23 77<br />

www.sika.com<br />

Innovation &<br />

Consistency<br />

since<br />

1910


Contents 4/<strong>2010</strong><br />

Information on <strong>SEI</strong><br />

www.iabse.org/sei<br />

<strong>SEI</strong> Advisory Board<br />

J.-E. Breen, W.F. Chen, Y. Fujino,<br />

N.J. Gimsing, P.R. Head, M. A. Hirt,<br />

D.A. Nethercot, M.J. Priestley,<br />

J. Schlaich, P. Taylor, M. Virlogeux,<br />

J.C. Walraven, H.F. Xiang.<br />

<strong>SEI</strong> Editorial Board<br />

H.H. Snijder, Chair; A. Schumacher,<br />

Vice Chair; M.G. Bruschi, A. Frangi,<br />

N.P. Hoej, K. Sugiura, D.Xu<br />

Correspondents<br />

China: D. Xu. Denmark: M. Braestrup.<br />

Egypt: F. Saad. Finland:<br />

M.-K. Söderqvist. France: B. Godart.<br />

Germany: U. Kuhlmann. India: V. Kumar.<br />

Italy: G. Bignotti, L. Ceriolo.<br />

Japan: K. Fujita. Korea: H.K. Kim.<br />

Norway: L. Toverud. Poland: W. Radomski.<br />

Russia: S.V. Mozalev. Sweden:<br />

H. Sundquist. Thailand: E. Limsuwan.<br />

UK: D.K. Doran. USA: J. Burns,<br />

D. Frangopol.<br />

Publisher<br />

IABSE<br />

ETH Zurich<br />

8093 Zurich, Switzerland<br />

Tel: 41-44-633 2647<br />

Fax: 41-44-633 1241<br />

secretariat@iabse.org<br />

www.iabse.org<br />

Publications Manager<br />

Brindarica Bose, IABSE<br />

Advertising Inquiries<br />

Sissel Niggeler, IABSE<br />

Published<br />

Quarterly: 1 Feb., 1 May, 1 Aug., 1 Nov.<br />

Subscription 2011<br />

Included in IABSE Membership.<br />

220 CHF: Individual Subscription<br />

630 CHF: Institutional Subscription<br />

Available through subscription agencies.<br />

ISSN 1016-8664, E-ISSN 1683-0350<br />

Copyright © IABSE. All rights reserved.<br />

Opinions and positions expressed in signed<br />

articles are those of the authors and are not<br />

necessarily those of Structural Engineering<br />

International or IABSE.<br />

Front cover:<br />

Wolchul Mountain Bridge, Korea<br />

See article on page 405<br />

Structural Engineering International<br />

International Association for Bridge and Structural Engineering<br />

Editorial<br />

Strengthening IABSE; P. L. Popovic; USA 361<br />

Special Feature: Fibre Reinforced Polymer Composites<br />

Scientific Papers*<br />

Introduction: Fiber Reinforced Polymer (FRP) Composites; A. Schumacher; Switzerland,<br />

M. D. G. Pulido; Spain 362<br />

Glass Fibre Reinforced Polymer Pultruded Flexural Members: Assessment of Existing<br />

Design Methods; J. R. Correia, F. Branco, J. Gonilha, N. Silva, D. Camotim; Portugal 362<br />

Effects of Hygrothermal Ageing on the Mechanical Properties of Glass-Fibre-Reinforced<br />

Polymer Pultruded Profiles; J. R. Correia, S. Cabral-Fonseca, A. Carreiro, R. Costa,<br />

M. P. Rodrigues, I. Eusébio, F. Branco; Portugal 370<br />

Evaluation of a Life Prediction Model and Environmental Effects of Fatigue for Glass Fiber<br />

Composite Materials; D. B. Dittenber, G. V. S. Hota; USA 379<br />

A Composite Bridge is Favoured by Quantifying Ecological Impact; R. A. Daniel; The Netherlands 385<br />

Experimental Assessment of Bond Behaviour of Fibre-Reinforced Polymers on Brick Masonry;<br />

E. Garbin, M. Panizza, M. Valluzzi; Italy 392<br />

Bridges with Glass Fibre–Reinforced Polymer Decks: The Road Bridge in Friedberg, Germany;<br />

J. Knippers, E. Pelke, M. Gabler, D. Berger; Germany 400<br />

Technical Reports<br />

Current and Future Applications of Glass-Fibre-Reinforced Polymer Decks in Korea; S. W. Lee,<br />

K. J. Hong, S. Park; Korea 405<br />

Field Issues Associated with the Use of Fiber-Reinforced Polymer Composite Bridge Decks<br />

and Superstructures in Harsh Environments; L. N. Triandafilou, J. S. O’Connor; USA 409<br />

Examples of Applications of Fibre Reinforced Plastic Materials in Infrastructure in Spain;<br />

A. Bansal, J. F. M. Cano, B. O. O. Muñoz, C. Paulotto; Spain 414<br />

Fiber-Reinforced Polymer Decks for Movable Bridges; R. D. Bottenberg; USA 418<br />

Glass Fiber Reinforced Polymer Strengthening and Evaluation of Railroad Bridge Members;<br />

G. V. S. Hota, P. V. Vijay, R. S. Abhari; USA 423<br />

Design of the St Austell Fibre-Reinforced Polymer Footbridge, UK; J. Shave, S. Denton,<br />

I. Frostick; UK 427<br />

General<br />

Scientific Papers*<br />

Aluminium Structures in Building and Civil Engineering Applications; F. Soetens; The Netherlands 430<br />

Glass Tensegrity Trusses; M. Froli, L. Lani; Italy 436<br />

A Simplified Serviceability Assessment of Footbridge Dynamic Behaviour Under Lateral<br />

Crowd Loading; L. Bruno, F. Venuti; Italy 442<br />

Technical Reports<br />

The Construction of the Main Bridge of the Yichang Yangtze River Railway Bridge in China;<br />

Y. Zhou, L. Zhang; China 447<br />

Static and Dynamic Analysis of the “Piedra Movediza” Replica Rock, Argentina; M. I. Montanaro,<br />

M. H. Peralta, N. Ercoli, M. L. Godoy, I. Rivas; Argentina 451<br />

Footbridge Studenci over the Drava River in Maribor, Slovenia; V. Markelj; Slovenia 454<br />

Sanchaji Bridge: Three-Span Self-Anchored Suspension Bridge, China; G. Dai, X. Song, N. Hu; China 458<br />

The First Extradosed Bridge in Slovenia; V. Markelj; Slovenia 462<br />

Recent PhD Abstracts 468<br />

Eminent Structural Engineer<br />

Christian Menn—Bridge Designer and Builder; E. Brühwiler; Switzerland 470<br />

Panorama<br />

IABSE Annual Meetings 473<br />

Predrag (Pete) Popovic, USA, New President of IABSE 473<br />

IABSE Awards <strong>2010</strong> 474<br />

The IABSE Foundation Anton Tedesko Medal 479<br />

IABSE Symposium Venice, September 22–24, <strong>2010</strong> 479<br />

Dhaka Conference ‘Advances in Bridge Engineering-II’ 482<br />

Calendar of Events and IABSE Members’ Business Cards 483<br />

IABSE Membership Application Form 484<br />

*Peer-reviewed papers<br />

Abstracting and Indexing: This publication is abstracted in Cambridge Scientific Abstracts under CSA Civil<br />

Engineering Abstracts; Emerald Abstracts; Construction and Building Abstracts (CBA); CAB Abstracts; INSPEC;<br />

and is included in EBSCOhost and SwetsWise Online Content. For <strong>SEI</strong> content Photocopying, Electronic usage, in<br />

the USA: Contact Copyrights Clearance Centre (CCC) at www.copyrights.com<br />

In rest of the world: Contact IABSE, at secretariat@iabse.org<br />

Structural Engineering International 4/<strong>2010</strong> 359


Building Asia together<br />

Holcim is building the very foundations of modern life and as a leading supplier of building<br />

materials in Asia we are strongly committed to the region.<br />

Global expertise and know-how, local market excellence and can-do attitude provide the<br />

strongest foundation for future growth. As with the Phu My Bridge, opened in September<br />

2009, that brought long awaited relief to Ho Chi Minh City, Vietnam.<br />

That’s what it takes to build with confidence in the most dynamic region in the world. We do<br />

this with respect for both the environment and the local communities where we operate.<br />

www.holcim.com<br />

Strength. Performance. Passion.


Strengthening IABSE<br />

This issue of Structural Engineering International contains a series of articles<br />

on strengthening structures with fiber-reinforced polymers (FRP). It seems<br />

appropriate that this Editorial should also be about strengthening—in particular<br />

strengthening IABSE.<br />

The Executive Committee under the leadership of Jacques Combault has been<br />

working to update IABSE’s long-range plan and is expecting to complete its work<br />

by the end of <strong>2010</strong>. The plan will outline strategies to strengthen IABSE.<br />

IABSE offers a well-established network for the comprehensive dissemination of<br />

knowledge to structural engineers worldwide through excellent conferences and<br />

publications such as Structural Engineering International (<strong>SEI</strong>) and Structural<br />

Engineering Documents (SED). Technical Commissions, Working Groups, and<br />

E-learning present additional platforms for exchange of technical information<br />

and ideas. However, despite these strengths, IABSE membership has not grown.<br />

Currently, IABSE has about 3600 members in 100 countries.<br />

There are several obvious reasons why our membership level has plateaued,<br />

including: competition from national technical societies and international societies<br />

focused on particular structure types or materials; lack of awareness among potential<br />

members of IABSE and the benefits it offers; high costs to attend conferences<br />

and pay membership fees for many members from developing countries; limited<br />

number of young engineers and students joining IABSE; and of course, the current<br />

world economic crisis causing a decrease in membership renewals. This latter trend<br />

could also have an impact over time on IABSE finances.<br />

To spur membership growth and in effect “strengthen” IABSE, we need to increase<br />

the visibility and relevance of IABSE to structural engineers from around the world.<br />

Stronger IABSE finances will follow increased membership. Our goal should be to<br />

attract 1000 new IABSE members over the next three years or to increase the net<br />

membership by at least 500. This will require all of our involvement and special<br />

efforts by National Groups. Specific strategies to address these challenges will be<br />

communicated as our long term plan is finalized.<br />

I am optimistic that together we will, despite the current economic environment, be<br />

able to strengthen and improve our organization over the next several years and<br />

that we will enjoy this journey together.<br />

Predrag L. Popovic<br />

President, IABSE<br />

Structural Engineering International 4/<strong>2010</strong> Editorial 361


Introduction: Fiber Reinforced Polymer (FRP) Composites<br />

Fiber reinforced polymer (FRP) composites can be considered<br />

a new class of construction material when compared<br />

with classical materials such as steel, concrete, timber and<br />

masonry. The relatively recent and growing interest in FRP<br />

in the domain of structural engineering can be traced to its<br />

advantageous properties ranging from a very high strengthto-weight<br />

ratio, electromagnetic neutrality, excellent fatigue<br />

behaviour, to superior durability including corrosion resistance.<br />

These properties have, in turn, lead to a broad spectrum<br />

of application that can be divided into two general<br />

categories: all-FRP members or structures in new construction<br />

or in the replacement of existing structural elements,<br />

and FRP components in the repair and rehabilitation of<br />

damaged or deteriorating structures.<br />

Structural Engineering International received an overwhelming<br />

response from around the world to its call for<br />

papers on the topic of FRP structures and strengthening of<br />

structures using FRP. The number of abstracts submitted,<br />

and subsequent high-quality papers received, has prompted<br />

the extension of this Special Edition over two issues—the<br />

present issue, as well as the coming May 2011 issue. In this<br />

first issue, six Scientific Papers on topics including existing<br />

design method assessments for FRP members, durability,<br />

environmental and fatigue issues for glass fiber reinforced<br />

polymer composites (GFRP), ecological advantages of FRP<br />

as compared with other materials, bond issues related to<br />

the use of FRP in the strengthening of masonry structures,<br />

and GFRP decks for bridges are presented. The Scientific<br />

Papers are complemented by six Technical Reports ranging<br />

from descriptions on the innovative use of FRPs in bridge<br />

decks to the application of GFRP in the strengthening of<br />

rail road bridges.<br />

Dr. Ann Schumacher, Vice-Chair <strong>SEI</strong> Editorial Board,<br />

Swiss Institute for Steel Construction, Switzerland<br />

Prof. M. Dolores G. Pulido, Chair WG 2 - Fiber<br />

Reinforced Polymer (FRP) Structures,<br />

Spanish National Research Council – Instituto<br />

CC Eduardo Torroja, Spain<br />

Glass Fibre Reinforced Polymer Pultruded Flexural<br />

Members: Assessment of Existing Design Methods<br />

João R. Correia, Prof. Dr, Technical Univ. of Lisbon, Instituto Superior Técnico/ICIST, Civil Eng. and Architecture,<br />

Lisbon, Portugal; Fernando Branco, Prof. Dr, Technical Univ. of Lisbon, Instituto Superior Técnico/ICIST, Civil Eng. and<br />

Architecture, Lisbon, Portugal; José Gonilha, Civil Eng., Technical Univ. of Lisbon, Instituto Superior Técnico/ICIST,<br />

Civil Eng. and Architecture, Lisbon, Portugal; Nuno Silva, Civil Eng., Technical Univ. of Lisbon, Instituto Superior<br />

Técnico/ICIST, Civil Eng. and Architecture, Lisbon, Portugal; Dinar Camotim, Prof. Dr, Technical Univ. of Lisbon,<br />

Instituto Superior Técnico/ICIST, Civil Eng. and Architecture, Lisbon, Portugal. Contact: jcorreia@civil.ist.utl.pt<br />

Abstract<br />

Glass fibre reinforced polymer (GFRP)<br />

pultruded profiles are being increasingly<br />

used in bridge and building construction<br />

as an alternative to traditional<br />

materials because of their several<br />

favourable properties that include high<br />

strength, low self-weight, short installation<br />

times, low maintenance requirements<br />

and improved durability. In spite<br />

of these advantageous characteristics,<br />

there are some factors delaying the<br />

widespread use of GFRP pultruded<br />

profiles in civil infrastructure, one of<br />

which is the lack of widely accepted<br />

design codes. This paper presents the<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: February 19, <strong>2010</strong><br />

Paper accepted: July 28, <strong>2010</strong><br />

results of analytical, experimental and<br />

numerical investigations on the structural<br />

behaviour of GFRP pultruded<br />

profiles, the objective of which was to<br />

evaluate the relative accuracy of existing<br />

design methods. A survey of analytical<br />

formulae available for the design<br />

of GFRP pultruded flexural members<br />

at both service and ultimate limit states<br />

is first presented. Subsequently, results<br />

of a test programme carried out at<br />

Instituto Superior Técnico (IST) are<br />

briefly discussed—the experiments<br />

included material characterization tests<br />

and full-scale flexural tests on I-section<br />

simply supported beams and cantilevers.<br />

These tests allowed for the evaluation<br />

of the service behaviour of GFRP<br />

flexural members and some of their<br />

most relevant failure mechanisms and<br />

respective ultimate loads. Results from<br />

experimental tests are compared with<br />

those obtained from analytical formulae<br />

and numerical models in order to<br />

evaluate the relative accuracy of existing<br />

design methods.<br />

Keywords: GFRP pultruded profiles;<br />

service behaviour; local buckling;<br />

global buckling; design methods; analytical<br />

formulae; numerical models.<br />

Introduction<br />

The limited durability of structures<br />

made with traditional materials and<br />

their consequent rehabilitation costs,<br />

which have substantially increased in<br />

the past few years, have been promoting<br />

the development of new structural<br />

materials that are less prone to corrosion,<br />

lighter and easier to erect. In this<br />

context, in the last two decades, fibre<br />

reinforced polymer (FRP) materials<br />

in general, and glass fibre reinforced<br />

polymer (GFRP) pultruded profiles<br />

in particular, have found a growing<br />

number of applications in buildings<br />

362 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


and bridges, in both new constructions<br />

and rehabilitation of degraded<br />

infrastructures. 1–10<br />

GFRP pultruded profiles have great<br />

potential as structural materials, presenting<br />

several advantages over traditional<br />

materials because of their<br />

high strength-to-weight ratio, low<br />

self-weight, electromagnetic transparency,<br />

possibility of being produced<br />

with any cross section, ease of installation,<br />

low maintenance requirements<br />

and improved durability under aggressive<br />

environments. 11 The drawback,<br />

in addition to the initial costs, lack of<br />

competitiveness for mainstream applications<br />

and the concerns regarding<br />

their behaviour under fire, 12,13 is that<br />

there are still no generally accepted<br />

design codes or guidelines available<br />

for civil engineering practitioners. As<br />

a consequence, at present, most structural<br />

designs are based on manufacturers’<br />

design guides, often presented in<br />

a tabular format, which are sometimes<br />

incomplete and over-conservative.<br />

The Eurocomp Design Code and<br />

Handbook, 14 published in 1996, provides<br />

design recommendations for<br />

polymer composites in general, but<br />

this non-normative document does not<br />

specifically address pultruded members.<br />

In 2002, the European Committee<br />

for Standardization (CEN) released<br />

the EN 13706 standard, 15 a normative<br />

document that merely defines two<br />

classes of pultruded profiles (associated<br />

with minimum values of material<br />

properties), not providing any<br />

design guidance. In 2007, the Italian<br />

National Research Council published<br />

the first national design guidelines<br />

for structures made of pultruded profiles;<br />

16 however, these specifications<br />

are mandatory only in Italy. It is also<br />

worth mentioning that most textbooks<br />

on the mechanics of composite materials<br />

and composite structures refer to<br />

aerospace and mechanical engineering<br />

applications—with the exception<br />

of a recent publication by Bank, 17<br />

which provides a comprehensive set of<br />

design rules for FRP structures, written<br />

in a civil engineering format.<br />

Before a comprehensive and widely<br />

accepted set of design rules and recommendations<br />

can be established<br />

for the use of GFRP pultruded profiles,<br />

further research work is needed<br />

to obtain in-depth understanding<br />

of their structural behaviour and<br />

to provide additional validation for<br />

the design methods that have been<br />

proposed.<br />

This paper presents the results of analytical,<br />

experimental and numerical<br />

investigations on the structural behaviour<br />

of GFRP pultruded profiles, the<br />

objective of which was to evaluate the<br />

relative accuracy of existing design<br />

methods. A survey of analytical formulae<br />

that have been suggested for<br />

the design of GFRP pultruded flexural<br />

members, for both service and<br />

ultimate limit states, is first presented.<br />

Subsequently, results of a test programme<br />

carried out at IST are briefly<br />

discussed—the experiments included<br />

material characterization tests on<br />

small-scale coupons and full-scale flexural<br />

tests on I-sections of simply supported<br />

GFRP beams and cantilevers.<br />

These tests, which are described in<br />

detail in Refs. [18, 19], allowed evaluation<br />

of the service behaviour of GFRP<br />

flexural members and some of their<br />

most relevant failure mechanisms and<br />

respective ultimate loads. The results<br />

from these experimental tests are then<br />

compared with predictions obtained<br />

from both analytical formulae and<br />

numerical models, in order to evaluate<br />

the relative accuracy of existing design<br />

methods.<br />

Design Methods for GFRP<br />

Flexural Members<br />

The design of structures made of<br />

GFRP pultruded profiles can be performed<br />

in much the same way as that<br />

of steel structures, provided that some<br />

necessary adaptations are taken into<br />

account, the most important of which<br />

are the orthotropic nature and linear<br />

elastic behaviour of the GFRP<br />

material.<br />

Thereafter, the structural design of<br />

standard GFRP profiles can be performed<br />

on the basis of either analytical<br />

beam models or shell and/or solid<br />

finite element (FE) models. For the<br />

former approach, which is most currently<br />

used in the design of GFRP<br />

frames and trusses, a simplified equivalent<br />

isotropic behaviour is assumed.<br />

For the latter approach, the orthotropic<br />

nature of the GFRP material is<br />

explicitly taken into account.<br />

Serviceability Limit States<br />

For service limit states design, the<br />

bending deflections of pultruded<br />

flexural members can be determined<br />

with a reasonable accuracy using<br />

analytical beam models based on the<br />

Timoshenko beam theory, that is,<br />

taking into account the shear contribution<br />

to overall deformation. In fact,<br />

shear deformations can be relatively<br />

important owing to the high elasticto-shear<br />

moduli ratio. For example,<br />

the elastic short-term deflection of a<br />

simply supported beam with a point<br />

load at midspan (similar to the beams<br />

whose experiments are reported later)<br />

can be calculated using Eq. (1),<br />

3<br />

P⋅L<br />

P⋅L<br />

δ = +<br />

48⋅E<br />

⋅I<br />

4⋅G<br />

⋅A<br />

full<br />

x<br />

full<br />

w<br />

(1)<br />

where d is the midspan deflection, P<br />

is the applied load, L is the span, I x is<br />

the second moment of area about the<br />

strong axis x, A w is the web(s) cross section<br />

and E full and G full are the full-scale<br />

longitudinal elastic and shear moduli<br />

for an equivalent isotropic behaviour,<br />

which can be determined on the basis<br />

of experiments (see next section).<br />

In order to evaluate long-term deflections<br />

in pultruded beams, it is necessary<br />

to address the viscoelastic response<br />

associated with the polymeric nature<br />

of the matrix properly. Therefore,<br />

time-dependent deformations due to<br />

sustained loads must be calculated taking<br />

into account the viscoelastic values<br />

of the full-scale moduli in Eq. (1). In<br />

Ref. [17], Bank presents a set of creep<br />

moduli and creep rate exponents recommended<br />

for design, which were<br />

obtained from long-term creep tests<br />

using the linearized version of Findley’s<br />

creep theory.<br />

Ultimate Limit States<br />

For ultimate limit states design, the fact<br />

that GFRP flexural members can theoretically<br />

collapse due to several failure<br />

modes must be taken into account. For<br />

the most commonly produced geometries<br />

(thin-walled open sections),<br />

the following failure mechanisms can<br />

occur: (a) flexural (tensile or compressive)<br />

failure; (b) web shear failure; (c)<br />

web transverse crushing; (d) local buckling;<br />

and (e) lateral-torsional buckling.<br />

Flexural Failure<br />

The bending moment associated with<br />

flexural failure of a pultruded member<br />

(M u ) can be calculated using Eq. (2),<br />

Mu = σ<br />

x , u⋅W<br />

(2)<br />

x<br />

where s x,u is the longitudinal failure<br />

stress (either compressive or tensile)<br />

of the GFRP material and W x is the<br />

cross-section elastic modulus about the<br />

strong axis. It is worth mentioning that<br />

flexural failure, due to compressive<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 363


crushing or tensile rupture, is not<br />

likely to occur for most common pultruded<br />

shapes, unless local buckling is<br />

prevented by an adequate stiffening<br />

system. 17<br />

Shear Failure<br />

The critical shear force (V u ) of a pultruded<br />

flexural member can be calculated<br />

using Eq. (3):<br />

V<br />

u<br />

τ<br />

u⋅I<br />

=<br />

S<br />

x<br />

x<br />

⋅t<br />

≈τ ⋅A<br />

(3)<br />

u<br />

where t u is the in-plane shear strength<br />

of the pultruded material, S x is the first<br />

moment of area about the strong axis,<br />

t is the laminate (web/flanges) thickness<br />

and A v is the shear area which,<br />

for most common profiles, corresponds<br />

to the web(s) of the profile. It should<br />

be noted that, similar to flexural failure,<br />

shear failure of the web material<br />

due to in-plane shear stresses seldom<br />

occurs, 17 as the strength of current<br />

cross sections and spans is dominated<br />

by buckling phenomena.<br />

Web Transverse Crushing<br />

The web(s) of GFRP pultruded<br />

beams can fail because of transverse<br />

crushing basically at two locations:<br />

(a) in the supports and (b) under concentrated<br />

loads. The critical crushing<br />

force (F crush u ) can be determined using<br />

Eq. (4):<br />

F<br />

crush<br />

u<br />

v<br />

c<br />

≈σ ,<br />

⋅A<br />

(4)<br />

y u<br />

eff<br />

where s c y,u can be taken as the transverse<br />

compressive strength and A eff is<br />

the effective cross section of the web<br />

subjected to the concentrated load,<br />

that is, the area of the web directly<br />

subjected to the support reaction or<br />

concentrated load. In order to avoid<br />

this failure mechanism, the lengths of<br />

the supports or loading patches can be<br />

increased and, in addition, web stiffeners<br />

can be used.<br />

Local Buckling due to In-Plane<br />

Compression<br />

Local buckling is an instability phenomenon<br />

characterized by transverse<br />

(flexural) bending of the member<br />

walls while the axis remains basically<br />

undeformed. Besides the high widthto-thickness<br />

ratios typically exhibited<br />

by thin-walled members made of any<br />

material (e.g. steel), GFRP pultruded<br />

profiles exhibit an added susceptibility<br />

to local buckling because of their<br />

reduced in-plane moduli. For the most<br />

common doubly symmetric profiles,<br />

the critical local buckling stress of<br />

flanges under compression can be<br />

determined using two alternative<br />

design formulae, derived by Kollár 20<br />

and by Mottram 21 —Eqs. (5) and (6),<br />

respectively,<br />

(Kollár 20 )<br />

local<br />

σ cr<br />

=<br />

( )<br />

2<br />

b / 2 ⋅<br />

⎛<br />

⎜<br />

⎝<br />

f<br />

1<br />

t<br />

L<br />

7 1 412<br />

f<br />

+ , ⋅<br />

×<br />

D ⋅ D<br />

T<br />

ξ I-flange<br />

⎞<br />

+ 12 ⋅D S⎟<br />

(5)<br />

⎠<br />

local<br />

where σ cr<br />

is the critical local buckling<br />

stress, b f is the flange width, t f is<br />

the flange thickness, D L , D T and D S<br />

are the longitudinal, transverse and<br />

shear flexural rigidities of the flange<br />

plate and x I-flange is the coefficient of<br />

edge restraint (assuming the flange is<br />

the critical wall); (Mottram 21 )<br />

σ<br />

local<br />

cr<br />

2 2<br />

π ⋅tf<br />

= ×<br />

2<br />

( b / 2<br />

f )<br />

⎡<br />

2<br />

⎛ b ⎞ E ⋅ E ⎤<br />

f<br />

L T<br />

⎢⎜<br />

045 , +<br />

2 ⎟<br />

⎥<br />

⎢⎝<br />

4a<br />

⎠ 12( 1−ν ⋅ν<br />

⎣<br />

L T)<br />

⎥<br />

⎦<br />

(6)<br />

where E L and E T are the in-plane<br />

longitudinal and transverse moduli,<br />

n L and n T are the major and minor<br />

Poisson’s ratios and a is the length of<br />

the buckle half-wavelength which, for<br />

I-section profiles, is suggested 21 to be<br />

taken as 3b f .<br />

With regard to the above-mentioned<br />

formulae, it should be mentioned<br />

that the use of Kollar’s design equations<br />

involves knowing all the in-plane<br />

properties (including the in-plane<br />

shear modulus) of both the web(s)<br />

and the flanges. Mottram’s alternative<br />

simplified procedure makes use of the<br />

flange’s properties only.<br />

Lateral-Torsional Buckling<br />

The critical lateral-torsional buckling<br />

stress for homogeneous doubly symmetric<br />

open profiles can be determined<br />

using the well-known Eurocode<br />

3 equation, adapted to the GFRP<br />

material orthotropy—Eq. (7):<br />

global<br />

Cb<br />

σ = ×<br />

cr<br />

Sx<br />

2<br />

4 2<br />

π ⋅E ⋅I ⋅G ⋅J<br />

π ⋅ E ⋅I<br />

⋅C<br />

L y LT<br />

L y w<br />

+<br />

2<br />

2 2<br />

( k ⋅ L<br />

f b)<br />

( k ⋅ L ) ( k ⋅ L<br />

f b w b)<br />

(7)<br />

where C b is a coefficient accounting<br />

for moment variation along the<br />

beam length, S x is the section modulus<br />

about the strong axis, I y is the second<br />

moment of area about the weak axis,<br />

J is the torsional constant, C w is the<br />

warping constant, k f is the effective<br />

length coefficient for flexural buckling<br />

about the weak axis (k f = 1,0 for simply<br />

supported beams, such as those tested<br />

in the experiments reported herein),<br />

k w is the effective length coefficient<br />

for torsional buckling of the section (in<br />

general, for simply supported beams,<br />

k w can be taken as 1,0) and L b is the<br />

unbraced length of the beam.<br />

For the particular case of cantilevers<br />

loaded at the shear centre of<br />

their extremity section, the critical<br />

lateral-torsional buckling load can be<br />

predicted using design formulae proposed<br />

by Timoshenko and Gere, 22<br />

assuming no warping at the fixed end<br />

and adapted to the GFRP material<br />

orthotropy—Eq. (8):<br />

EIG J<br />

global L y LT<br />

Pcr<br />

= γ 2<br />

(8)<br />

2<br />

L<br />

where P cr is the critical lateral-torsional<br />

buckling load, and g 2 is a dimensionless<br />

factor depending on the torsional<br />

and warping rigidities.<br />

Experimental Assessment of<br />

the Design Methods<br />

As already mentioned, the establishment<br />

of consensual design approaches<br />

is dependent on further validation of<br />

the existing design methods and, most<br />

likely, on the development of new<br />

methodologies. In order to contribute<br />

to achieving this goal, a research effort<br />

was conducted at IST, which consisted<br />

of a fairly extensive experimental<br />

investigation, described in detail<br />

in Refs. [18, 19], whose results were<br />

then compared with several different<br />

types of numerical simulations. 23<br />

In this study, the experimental results<br />

obtained in the above investigation<br />

are used to assess the accuracy of the<br />

design methods described earlier.<br />

The experimental investigation<br />

involved pultruded GFRP I-beams (a)<br />

made of an isophthalic polyester matrix<br />

reinforced with E-glass fibre rovings<br />

and mats (inorganic content of 62%,<br />

by weight) and (b) exhibiting the following<br />

nominal dimensions: web height<br />

of 200 mm, flange width of 100 mm<br />

and thickness of 10 mm. The experimental<br />

study comprised (a) material<br />

characterization tests, to evaluate the<br />

mechanical properties and response of<br />

the GFRP material; (b) flexural tests<br />

on simply supported beams, aimed at<br />

364 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


evaluating their behaviour under service<br />

and failure conditions (including<br />

web transverse crushing and local and<br />

lateral-torsional buckling); and (c) flexural<br />

tests on cantilevers to investigate<br />

their lateral-torsional buckling behaviour<br />

and failure under tip-point loads<br />

applied at different locations. Each test<br />

type is addressed individually in the<br />

following sections. After providing the<br />

experimental set-up and procedure, the<br />

relevant results are outlined and used to<br />

assess the quality (accuracy and safety)<br />

of the corresponding design method.<br />

Material Characterization Tests<br />

Interlaminar shear tests (ASTM<br />

D2344) were first conducted on<br />

specimens with nominal dimensions<br />

of 9,8 × 20,0 × 60,0 mm 3 , applying a<br />

concentrated load at the centre of a<br />

45,0 mm span, in order to determine<br />

the interlaminar shear strength (F sbs ).<br />

Three-point bending tests (ISO 14125)<br />

were then performed on specimens<br />

with nominal dimensions of 9,8 × 15,0<br />

× 300 mm 3 , tested in the longitudinal<br />

direction (L), in order to determine<br />

the flexural strength (s fu,L ), the elastic<br />

modulus in bending (E f,L ) and the<br />

strain at failure (e fu,L ). Tensile tests<br />

(ISO 527-1,4) were also performed,<br />

using specimens with nominal dimensions<br />

of 9,8 × 15,0 × 350 mm 3 , loaded<br />

in their longitudinal direction, allowing<br />

measurement of the tensile strength<br />

(s tu,L ), the strain at failure (e tu,L ), the<br />

elastic modulus in tension (E t,L ) and<br />

the Poisson’s ratio (n LT ). Finally, compressive<br />

tests (ASTM D695) were carried<br />

out on specimens with nominal<br />

dimensions of 9,8 × 12,7 × 39,0 mm 3 ,<br />

in order to determine, for both longitudinal<br />

(L) and transverse (T) directions,<br />

the compressive strength (s cu,L and<br />

s cu,T ), the strain at failure (e cu,L and<br />

e cu,T ) and the elastic modulus in compression<br />

(E c,L and E c,T ).<br />

In all mechanical tests, the material<br />

generally exhibited linear-elastic<br />

behaviour until failure, a typical feature<br />

of the GFRP material. 18,19 The<br />

failure modes observed in the different<br />

mechanical tests are illustrated in<br />

Fig. 1. Table 1 presents a summary of<br />

the mechanical properties obtained in<br />

these tests (which will be later used<br />

as input data in the analytical and<br />

numerical design methods), namely,<br />

the ultimate stress (s u ), the elastic<br />

modulus (E), the strain at failure (e u ),<br />

the Poisson’s ratio (n LT ), the interlaminar<br />

shear strength (F sbs ) and the<br />

in-plane shear strength (τ u ), the latter<br />

obtained from tensile tests on double<br />

lap bolted joints. 18 It is worth mentioning<br />

that the behaviour exhibited by<br />

coupons extracted from both the web<br />

and the flanges was similar. The different<br />

mechanical properties in tension,<br />

flexure and compression, together<br />

with the material orthotropy, are also<br />

outlined.<br />

Property/test<br />

and direction<br />

Longitudinal<br />

flexure<br />

Flexural Behaviour of Simply<br />

Supported Beams<br />

Test Set-up and Results<br />

This experimental series consisted of<br />

four beams with different spans and<br />

lateral bracing systems, all subjected to<br />

a point load at midspan. Beams V1 and<br />

V2 were both tested in a 4,00 m span,<br />

while beams V3 and V4 were tested in<br />

spans of 1,44 and 1,00 m, respectively.<br />

In beam V1, in order to prevent lateraltorsional<br />

instability, a lateral bracing<br />

system was used along the beam span<br />

(Fig. 2). All the other beams were laterally<br />

unrestrained. Load was applied<br />

using a hydraulic jack that transmitted<br />

the load to the top flange of the GFRP<br />

profile through square steel spreading<br />

plates of side 0,08 m; a metallic sphere<br />

was placed between the two spreading<br />

plates, in order to avoid any transverse<br />

loading. In beam V2, in order to<br />

investigate the influence of the loading<br />

system in restraining the beam at midspan,<br />

successive changes were introduced<br />

in the loading system. 18,19 The<br />

supports of all the beams were made<br />

of 0,05 m diameter steel rollers with<br />

L ongitudinal<br />

tension<br />

Longitudinal<br />

compression<br />

Transverse<br />

compression<br />

s u (MPa) 624,6 ± 26,9 475,5 ± 25,5 375,8 ± 67,9 122,0 ± 15,4<br />

E (GPa) 26,9 ± 1,3 32,8 ± 0,9 26,4 ± 1,9 7,4 ± 0,4<br />

e u (10 −3 ) 24,9 ± 1,3 15,4 ± 1,5 17,0 ± 2,5 21,5 ± 1,7<br />

n xy (–) — 0,28 — —<br />

Interlaminar shear strength, F sbs = 35,0 ± 3,9 MPa.<br />

In-plane shear strength, t u = 38,7 ± 5,6 MPa.<br />

Full-scale properties (equivalent isotropic behaviour): E full = 38,3 GPa; G full = 3,58 GPa.<br />

Table 1: Mechanical properties of the GFRP profile from coupon (average and standard<br />

deviation values) and full-scale testing<br />

(a)<br />

(b) (c) (d)<br />

Fig. 1: Failure modes: (a) interlaminar shear; (b) flexure; (c) tension; and (d) compression<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 365


Fig. 2: Beam V1—deformation on the brink<br />

of collapse<br />

0,08 m long top steel plates; both end<br />

supports allowed for free rotation and<br />

one of them also allowed for longitudinal<br />

sliding.<br />

All beams presented linear-elastic<br />

behaviour up to failure. 18,19 The fullscale<br />

elastic constants of the GFRP<br />

profile were estimated on the basis of<br />

the method proposed by Bank, 24 which<br />

involves performing a linear regression<br />

analysis of the slope of the load-deflection<br />

curves for varying spans. This<br />

analysis provided a longitudinal elastic<br />

modulus (E full ) of 38,3 GPa and a shear<br />

modulus (G full ) of 3,58 GPa. One can<br />

readily note that the elastic constants<br />

provided by mechanical tests on smallscale<br />

specimens (E t,x = 32,8 GPa; E f,x =<br />

26,9 GPa, cf. Table 1) may differ considerably<br />

from those obtained in fullscale<br />

tests. Such variation is mainly due<br />

to the inhomogeneous constitution<br />

of both the GFRP laminates and the<br />

overall cross section and also to differences<br />

in the experimental set-up.<br />

Failure of beam V1 occurred due to<br />

local buckling of the top flange, for<br />

a midspan deflection of 107,3 mm<br />

(about 1/37 of the span, Fig. 2) and a<br />

load of 60,2 kN, which corresponded to<br />

a longitudinal maximum stress of 268,2<br />

MPa. Failure occurred with delamination<br />

of the top flange and web-top<br />

flange separation in the vicinity of<br />

midspan (Fig. 3), followed by web<br />

transverse bending. This test showed<br />

the importance of the local buckling<br />

phenomenon in members under compression,<br />

such as the flanges of bended<br />

beams. In fact, at failure, the maximum<br />

longitudinal stress was about 56 and<br />

71% of the tensile and compressive<br />

material strengths, respectively.<br />

In several iterations of the flexural<br />

test of beam V2, failure was always<br />

triggered by lateral-torsional buckling<br />

(Fig. 4). The different test set-ups<br />

proved to have a significant influence<br />

on the buckling load, which varied<br />

from 13,0 to 20,7 kN. The lowest buckling<br />

load (minimum restriction introduced<br />

by the load application system at<br />

midspan) corresponded to a maximum<br />

longitudinal stress of 58,0 MPa, showing<br />

the importance of global instability<br />

in slender unrestrained beams.<br />

Failure of beams V3 and V4 was due<br />

to crushing of the web at midspan<br />

(Fig. 5) under the applied load,<br />

and occurred for loads of 88,2 and<br />

107,5 kN, respectively. Crushing failure<br />

of the web was followed by the<br />

development of longitudinal cracks<br />

in the web–top flange junction. The<br />

above-mentioned failure loads correspond<br />

to maximum transverse compressive<br />

stresses in the web (under the<br />

applied load) of 112,6 and 137,1 MPa,<br />

calculated using Eq. (4).<br />

Assessment of Design Methods<br />

In addition to the analytical formulae<br />

presented earlier, FE models of all<br />

Fig. 4: Beam V2—lateral-torsional buckling<br />

tested GFRP beams were developed<br />

using the FE program SAP2000. 25 The<br />

web and flanges of the GFRP profiles<br />

were modelled using adapted DKQ<br />

(discrete Kirchhoff quadrilateral)<br />

shell elements, consisting of four-node<br />

rectangular/triangular elements with<br />

bilinear interpolation functions. 25,26<br />

Linear-elastic orthotropic material<br />

behaviour was assumed on the basis<br />

of the results of experiments (cf.<br />

Table 1). The actual supporting conditions<br />

were simulated with node constraints.<br />

Both linear static and linear<br />

buckling analyses were carried out—<br />

Fig. 6 shows the buckled configuration<br />

of beam V1.<br />

For the serviceability behaviour, the<br />

deflections of all tested beams could<br />

be back-calculated with a very high<br />

accuracy (maximum and average relative<br />

errors among the tested beams of<br />

3 and 1%, respectively) on the basis of<br />

Timoshenko beam theory and using<br />

the calculated full-scale elastic constants<br />

in Eq. (1). It is also worth mentioning<br />

that in all tested beams, the<br />

shear contribution to deformation was<br />

significant: 12,6% for the relatively<br />

slender beams V1 and V2; 41,1 and<br />

59,1% for the less slender beams V3<br />

and V4, respectively.<br />

With regard to the failure behaviour,<br />

the experimental critical load of beam<br />

V1 (60,2 kN) compared reasonably<br />

well with numerical (53,9 kN, relative<br />

difference of −10,5%) and analytical<br />

predictions, the latter obtained using<br />

the two alternative design formulae<br />

presented by Kollár 20 (58,0 kN, −3,7%)<br />

and Mottram 21 (56,8 kN, −5,6%).<br />

Analytical predictions were computed<br />

on the basis of Eqs. (5) and (6), respectively,<br />

using the coupon material properties<br />

(cf. Table 1); in these calculations,<br />

a standard value of ν T = 0,10 was considered<br />

and it was assumed that the<br />

in-plane shear modulus, G LT , is equal<br />

to the full-scale shear modulus, G full ).<br />

The differences between experimental<br />

and predicted critical loads are very<br />

Fig. 3: Beam V1—local buckling failure<br />

Fig. 5: Beam V3—web crushing under applied load<br />

366 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


(Fig. 8)—firstly, in the curves corresponding<br />

to the horizontal deflection of<br />

both flanges (d 2 and d 3 ) and secondly,<br />

in the curve describing the vertical<br />

deflection of the shear centre (d 1 ).<br />

As expected, the critical bucking load<br />

decreased with increasing span and, for<br />

each span the highest critical load was<br />

obtained when the load was applied<br />

at the centre of the bottom flange<br />

(BF), while the lowest critical load was<br />

obtained when loading at the centre<br />

of the top flange (TF), as illustrated in<br />

Fig. 9. For the shortest span of 2,0 m,<br />

maximum longitudinal stresses varied<br />

between 53,3 and 125,0 MPa, while for<br />

the longest span of 4,0 m those stresses<br />

varied between 24,7 and 44,7 MPa.<br />

Fig. 6: FE model of beam V1—local buckling configuration<br />

small, particularly if the relatively high<br />

coefficients of variation exhibited by<br />

GFRP material properties are taken<br />

into account. In principle, because of<br />

the geometric imperfections of the<br />

material, one would expect the experimental<br />

results to be below analytical/<br />

numerical predictions; the fact that<br />

predictions are slightly lower than<br />

the experimental critical load, has to<br />

be attributed to the above-mentioned<br />

material inhomogeneity and, eventually,<br />

to some slight restriction introduced<br />

by the loading system.<br />

For beam V2, the minimum experimental<br />

critical load (13,0 kN) differed<br />

quite significantly from numerical<br />

(5,0 kN) and analytical (4,7 kN) predictions,<br />

obtained respectively with the<br />

above-mentioned FE model and using<br />

Eq. (7), again computed using coupon<br />

material properties (cf. Table 1). For<br />

this beam, the restriction (friction)<br />

introduced by the loading system had<br />

a very significant effect in preventing<br />

the triggering of lateral-torsional buckling.<br />

Therefore, in order to understand<br />

this instability mechanism better and<br />

to assess the accuracy of analytical and<br />

numerical design tools, it was decided<br />

to perform tests on cantilevers, for<br />

which it is easier to prevent the loading<br />

system from restraining deformations<br />

(see next section).<br />

For beams V3 and V4, comparison of<br />

the maximum transverse compressive<br />

stresses in the web under the applied<br />

load (112,6 and 137,1 MPa, respectively,<br />

obtained by dividing the applied<br />

load by the area of the web directly<br />

under the metal plates positioned<br />

below the hydraulic jacks) with the<br />

compressive strength in the transverse<br />

direction, obtained from the material<br />

characterization tests (122,0 MPa, cf.<br />

Table 1), allows justifying the observed<br />

failure mode.<br />

Flexural Behaviour of Cantilevers<br />

Test Set-up and Results<br />

In this experimental series, three different<br />

spans of 2,0, 3,0 and 4,0 m were<br />

tested and, for each span, the load was<br />

applied in three alternative positions<br />

of the free end cross section: at the<br />

centre of the top flange (TF), at the<br />

centroid or shear centre (SC) and at<br />

the centre of the bottom flange (BF).<br />

Load was applied using a dead-load<br />

system, consisting of a metal bucket<br />

filled with metal plates and water,<br />

which was suspended from the free end<br />

cross section of the GFRP cantilevers,<br />

at the three predefined positions. The<br />

vertical support of the cantilevers was<br />

made of a thick steel plate connected<br />

to a transverse metal beam, which was<br />

placed over the top flange of the profile<br />

using four Dywidag bars. Horizontal<br />

deflections in the support section were<br />

restrained by means of metallic plates<br />

and sets of metallic bolts placed on<br />

both sides of the web.<br />

All cantilevers tested presented lateraltorsional<br />

buckling—Fig. 7 illustrates the<br />

buckled configuration of a 4,0 m span<br />

cantilever loaded at the SC. This global<br />

instability could easily be distinguished<br />

in the load-deflection behaviour<br />

Fig. 7: Lateral-torsional buckling of a 4,0 m<br />

span cantilever (load at SC)<br />

Load (kN)<br />

2,25<br />

2,00<br />

1,75<br />

1,50<br />

1,25<br />

1,00<br />

d 3<br />

0,75<br />

0,50<br />

d 1<br />

0,25<br />

d 2<br />

0,00<br />

0 20 40 60 80<br />

Deflection (mm)<br />

Fig. 8: Load-deflection curves for a 4,0<br />

span cantilever (load at SC)<br />

Critical load (kN)<br />

Fig. 9: Critical load as a function of the<br />

span, for different load positions<br />

100<br />

18<br />

16<br />

BF — experiment<br />

SC — experiment<br />

14<br />

TF — experiment<br />

12<br />

10<br />

8<br />

SC — analytical<br />

BF — numerical<br />

SC — numerical<br />

TF — numerical<br />

6<br />

4<br />

2<br />

0<br />

1,5 2,0 2,5 3,0 3,5 4,0 4,5<br />

Cantilever span (m)<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 367


supplying the GFRP profiles used in the<br />

experimental investigations.<br />

References<br />

Fig. 10: FE model of a 4,0 m span cantilever (load at TF) —global buckling configuration<br />

Assessment of Design Methods<br />

Shell FE models of all tested cantilevers,<br />

similar to those described for the simply<br />

supported beams, were developed<br />

with the program SAP2000. Figure 10<br />

illustrates the buckling mode for a<br />

4,0 m span cantilever loaded at the TF.<br />

Figure 9 shows the comparison<br />

between experimental critical loads<br />

and numerical predictions using the<br />

aforementioned FE models (for all<br />

three load positions) and analytical<br />

predictions (computed only for shear<br />

centre loading) using Eq. (8).<br />

Average relative errors for analytical<br />

and numerical predictions are 4,7<br />

and 12,9%, respectively. Therefore,<br />

it seems fair to claim a reasonable<br />

agreement between experimental critical<br />

loads and both design methods—<br />

predictions unarguably exhibit a similar<br />

pattern of variation with both the<br />

span and the load position. Figure 9<br />

also shows that, in general, experimental<br />

critical loads are lower than predictions,<br />

which may be attributed to the<br />

effect of geometrical imperfections in<br />

the profile and, in addition, to the not<br />

completely fixed restraint condition at<br />

the support.<br />

Conclusion<br />

The flexural behaviour of GFRP<br />

pultruded profiles presents several differences<br />

when compared to traditional<br />

materials, at both material and structural<br />

levels. On one hand, contrary to<br />

steel that yields and concrete that cracks,<br />

in general, GFRP profiles present<br />

linear-elastic behaviour until failure,<br />

which usually occurs with large deformations.<br />

On the other hand, the design<br />

of GFRP members is often governed<br />

by deformability restrictions, due to the<br />

low elastic modulus in the longitudinal<br />

direction and also due to the contribution<br />

of shear to the global deformation.<br />

In general, design at ultimate limit states<br />

is not governed by material strength, as<br />

failure is usually due to local or global<br />

buckling phenomena, with stress levels<br />

in service that are relatively low, in spite<br />

of the high strength exhibited by the<br />

GFRP material.<br />

The comparison between experimental<br />

results and predicted behaviour with the<br />

FE models and analytical formulae presented<br />

in this paper for a limited range<br />

of failure scenarios shows that numerical<br />

and analytical tools available for<br />

the design of GFRP pultruded flexural<br />

members are reasonably accurate. In<br />

particular, for service design, the bending<br />

deformations of GFRP beams can<br />

be readily calculated with Timoshenko’s<br />

beam theory, using full-scale elastic<br />

constants and assuming an equivalent<br />

isotropic behaviour. At ultimate limit<br />

states, failure mechanisms associated<br />

with material crushing and local or<br />

global buckling can be easily computed,<br />

using laminate material properties,<br />

obtained through coupon testing.<br />

Acknowledgements<br />

The authors wish to acknowledge the support<br />

of FCT, ICIST and Agência da Inovação<br />

(Grant No. 2009/003456) for funding the<br />

research and also STEP and ALTO for<br />

[1] Keller T. Recent all-composite and hybrid<br />

fibre-reinforced polymer bridges and buildings.<br />

Prog. Struct. Eng. Mater. 2001; 3(2): 132–140.<br />

[2] Keller T. Use of Fibre Reinforced Polymers<br />

in Bridge Construction. Structural Engineering<br />

Documents, vol. 7. IABSE: Zurich, 2003.<br />

[3] Burgoyne C. Advanced composites in civil<br />

engineering in Europe, Struct. Eng. Int. 1999;<br />

9(4): 267–273.<br />

[4] Braestrup M. Footbridge constructed from<br />

glass-fiber reinforced profiles, Denmark. Struct.<br />

Eng. Int. 1999; 9(4): 256–258.<br />

[5] GangaRao H, Craigo II CA. Fiber reinforced<br />

composite bridge decks in the USA. Struct. Eng.<br />

Int. 1999; 9(4): 286–288.<br />

[6] Keller T. Towards structural forms for composite<br />

fibre materials. Struct. Eng. Int. 1999; 9(4):<br />

297–300.<br />

[7] Sobrino JA, Pulido MDG. Towards advanced<br />

composite material footbridges. Struct. Eng. Int.<br />

2002; 12(2): 84–86.<br />

[8] Cheng L, Karbhari VM. New bridge systems<br />

using FRP composites and concrete: a state-ofthe-art<br />

review. Prog. Struct. Eng. Mater. 2006;<br />

8(4): 143–154.<br />

[9] Neto ABS, La Rovere HL. Composite concrete/GFRP<br />

slabs for footbridge deck systems.<br />

Compos. Struct. <strong>2010</strong>; 92(10): 2554–2564.<br />

[10] Hollaway LC. A review of the present and<br />

future utilisation of FRP composites in the civil<br />

infrastructure with reference to their important<br />

in-service properties. Const. Build. Mater., in<br />

press, doi:10.1016/j.conbuildmat.<strong>2010</strong>.04.062.<br />

[11] Correia JR, Cabral-Fonseca S, Branco<br />

FA, Ferreira J, Eusébio MI, Rodrigues MP.<br />

Durability of glass fibre reinforced polyester<br />

(GFRP) pultruded profiles for construction<br />

applications. Mech. Compos. Mater. 2006; 42(4):<br />

325–338.<br />

[12] Correia JR, Branco FA, Ferreira J, Cabral-<br />

Fonseca S, Rodrigues JPC. Lifetime performance<br />

of GFRP pultruded profiles for structural<br />

applications. IABSE Symposium Improving<br />

Infrastructure—Bringing People Closer Worldwide,<br />

Weimar, 2007.<br />

[13] Correia JR. GFRP Pultruded Profiles in<br />

Civil Engineering: Hybrid Solutions, Bonded<br />

Connections and Fire Behaviour, PhD Thesis,<br />

IST, Technical University of Lisbon, 2008.<br />

[14] Clarke JL (ed.). Structural Design of<br />

Polymer Composites—EuroComp Design Code<br />

and Handbook. E & FN Spon: London, 1996.<br />

[15] CEN. EN 13706: Reinforced Plastics<br />

Composites—Specifications for Pultruded<br />

Profiles. Part 1: Designation; Part 2: Methods of<br />

Test and General Requirements; Part 3: Specific<br />

Requirements. European Committee for<br />

Standardisation: Brussels, 2002.<br />

[16] National Research Council of Italy. Guide<br />

for the Design and Construction of Structures<br />

made of FRP Pultruded Elements, Advisory<br />

Committee on Technical Recommendations for<br />

Construction: Roma, 2008.<br />

368 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


[17] Bank LC. Composites for Construction:<br />

Structural Design with FRP Materials. Wiley:<br />

Hoboken, NJ, 2006.<br />

[18] Correia JR. GFRP Pultruded Profiles:<br />

The Use of GFRP-concrete Hybrid Beams<br />

in Construction, MSc Thesis, IST, Technical<br />

University of Lisbon, 2004 (in Portuguese).<br />

[19] Correia JR, Branco FA, Silva NMF,<br />

Camotim D, Silvestre N. First-order, buckling<br />

and post-buckling behaviour of GFRP pultruded<br />

beams—part 1: experimental study, submitted for<br />

publication. Contact author for details: jcorreia@<br />

civil.ist.utl.pt<br />

[20] Kollár LP. Local buckling of fiber reinforced<br />

plastic composite structural members with open<br />

and closed cross sections. J. Struct. Eng. 2003;<br />

129(11): 1503–1513.<br />

[21] Mottram JT. Determination of critical load<br />

for flange buckling in concentrically loaded pultruded<br />

columns. Compos. B: Eng. 2004; 35(1):<br />

35–47.<br />

[22] Timoshenko SP, Gere JM. Theory of Elastic<br />

Stability, McGraw-Hill: New York, 1963.<br />

[23] Silva NMF, Camotim D, Silvestre N, Correia<br />

JR, Branco FA. First-order, buckling and<br />

post-buckling behaviour of GFRP pultruded<br />

beams—part 2: numerical simulation, submitted<br />

for publication. Contact author for details:<br />

jcorreia@civil.ist.utl.pt<br />

[24] Bank LC. Flexural and shear moduli of fullsection<br />

fiber reinforced plastic (FRP) pultruded<br />

beams. J. Testing Eval. 1989; 17(1): 40–45.<br />

[25] Computers and Structures Inc. SAP 2000<br />

User’s Manual (version 11.0), 2007.<br />

[26] Zienkiewicz OC, Taylor RL. The finite<br />

element method—Volume 2. Solid and fluid<br />

mechanics. Dynamics and Non-Linearity.<br />

McGraw-Hill: London, 1991.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 369


Effects of Hygrothermal Ageing on the Mechanical<br />

Properties of Glass-Fibre-Reinforced Polymer Pultruded<br />

Profiles<br />

João R. Correia, Prof., Dr, Technical Univ. of Lisbon, Instituto Superior Técnico/ICIST, Civil Eng. and Architecture, Lisbon, Portugal;<br />

Susana Cabral-Fonseca, Dr, Eng., Laboratório Nacional de Engenharia Civil, Lisbon, Portugal; Ana Carreiro, Civil Eng., Technical<br />

Univ. of Lisbon, Instituto Superior Técnico, Civil Eng. and Architecture, Lisbon, Portugal; Ricardo Costa, Civil Eng., Technical Univ.<br />

of Lisbon, Instituto Superior Técnico, Civil Eng. and Architecture, Lisbon, Portugal; Maria Paula Rodrigues, Dr, Eng., Laboratório<br />

Nacional de Engenharia Civil, Lisbon, Portugal; Isabel Eusébio, Dr, Eng., Laboratório Nacional de Engenharia Civil, Lisbon, Portugal;<br />

Fernando Branco, Prof., Dr, Technical Univ. of Lisbon, Instituto Superior Técnico/ICIST, Civil Eng. and Architecture, Lisbon,<br />

Portugal. Contact: jcorreia@civil.ist.utl.pt<br />

Abstract<br />

This paper presents the results of an experimental study on the physical and<br />

mechanical changes suffered by glass-fibre-reinforced polymer (GFRP) pultruded<br />

profiles, made of either unsaturated polyester or vinylester resins, after<br />

accelerated hygrothermal ageing. Specimens from both types of profiles, comprising<br />

identical fibre contents and architectures, were subjected to: (a) immersion<br />

in demineralized water; (b) immersion in saltwater at temperatures of 20, 40<br />

and 60°C for 12 months and; (c) continuous condensation at 40°C for 9 months.<br />

Batches of test specimens from both profiles, conditioned in those accelerated<br />

exposure environments, were periodically monitored with respect to: (a) mass<br />

changes; (b) variation in glass transition temperature evaluated through dynamic<br />

mechanical analysis (DMA) and; (c) degradation of mechanical properties,<br />

assessed by means of tensile, flexural and interlaminar shear tests.<br />

Keywords: GFRP; unsaturated polyester matrix; vinylester matrix; pultruded<br />

profiles; hygrothermal ageing; mechanical properties.<br />

Introduction<br />

Fibre-reinforced polymer (FRP) materials<br />

in general, and glass-fibre-reinforced<br />

polymer (GFRP) pultruded<br />

profiles in particular, are being used<br />

increasingly in civil engineering applications<br />

as an alternative to traditional<br />

materials, such as steel, reinforced concrete<br />

and timber. This growing acceptance<br />

of FRP structures, particularly in<br />

corrosive applications, can be attributed<br />

to their improved durability and low<br />

maintenance requirements, in addition<br />

to other intrinsic advantageous properties<br />

of advanced composite materials<br />

that include high strength, lightness<br />

and low thermal conductivity. 1,2<br />

In regard to durability, the long-term<br />

use of FRP materials in vessels, pipelines,<br />

storage tanks and chemicalresistant<br />

equipment of the oil industry<br />

provides evidence of their improved<br />

performance in relatively harsh and<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: February 19, <strong>2010</strong><br />

Paper accepted: July 25, <strong>2010</strong><br />

corrosive environments, when compared<br />

to traditional materials. However,<br />

for civil engineering applications; owners,<br />

designers and contractors request<br />

comprehensive and validated data on<br />

durability, since the service life of mainstream<br />

structures is generally expected<br />

to exceed 50 years. As most FRP civil<br />

engineering structures are quite recent 3<br />

and research already carried out on this<br />

topic is still limited, such information<br />

correlating the effects of environmental<br />

degradation on the physical, chemical<br />

and mechanical properties of FRPs<br />

is currently not available. The development<br />

of reliable degradation models,<br />

similar to those already available for<br />

traditional materials, involves gathering<br />

such comprehensive data on durability.<br />

It is also worth mentioning that comparative<br />

studies on the performance of<br />

alternative matrix formulations used in<br />

FRP materials are also scarce. In this<br />

context, paradoxically, the widespread<br />

acceptance of FRP materials is delayed<br />

because of concerns about durability.<br />

In this regard, several authors have<br />

recently identified durability as one<br />

of the most critical gap between perceived<br />

need for information and available<br />

information, and as a crucial area<br />

of focus for future research on FRP<br />

materials. 4,5<br />

This paper presents results of an<br />

experimental study on the physical and<br />

mechanical changes suffered by GFRP<br />

pultruded profiles, made of either<br />

unsaturated polyester or vinylester<br />

resins with identical fibre contents and<br />

architectures, following accelerated<br />

hygrothermal ageing. Specimens from<br />

both types of profiles were subjected<br />

to immersion in demineralized water<br />

and saltwater for temperatures of 20,<br />

40 and 60°C and, in addition, to continuous<br />

condensation at 40°C, simulating<br />

the ageing conditions in wet<br />

environments (e.g. placed under water<br />

or subjected to high levels of moisture),<br />

in coastal areas or where the<br />

use of de-icing salts is common. Other<br />

environmental degradation agents,<br />

not investigated in the present study,<br />

include acid or alkaline fluids, thermal<br />

cycles, freeze–thaw cycles, ultraviolet<br />

radiation and elevated temperature. 4<br />

Batches of test specimens from both<br />

types of profiles, placed in the abovementioned<br />

degradation environments,<br />

were periodically removed and monitored<br />

regarding: (a) mass changes;<br />

(b) variation in glass transition temperature<br />

evaluated through dynamic<br />

mechanical analysis (DMA) and; (c)<br />

degradation of mechanical properties<br />

in tension, flexure and shear.<br />

This paper provides extensive data<br />

on the effects of hygrothermal ageing<br />

on the performance of GFRP pultruded<br />

profiles, thereby contributing<br />

to shortening of the above-mentioned<br />

gap between perceived need for information<br />

and available information on<br />

durability. In addition, as similar fibre<br />

contents/architectures were used in<br />

this study, results obtained allow for a<br />

direct comparison between the performances<br />

of unsaturated polyester and<br />

vinylester resins.<br />

370 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Experimental Programme<br />

Materials<br />

The material studied was obtained<br />

from two commercial GFRP pultruded<br />

tubular profiles (50 × 50 mm,<br />

thickness 5 mm). This material consists<br />

of alternating layers of unidirectional<br />

E-glass fibre rovings and strand mats<br />

embedded in either unsaturated polyester<br />

resin (profile “UP”) or vinylester<br />

resin (profile “VE”). The former resin<br />

is used in most structural applications<br />

when there are no particular requirements<br />

in terms of environmental<br />

aggressiveness, whereas the latter is<br />

often selected for applications in relatively<br />

harsh or corrosive environments.<br />

The two profiles were produced with<br />

the same glass fibre content and architecture<br />

(Figs. 1 and 2), thus allowing<br />

comparison of the durability performance<br />

of the polyester and vinylester<br />

resins used in these off-the-shelf standard<br />

profiles.<br />

Initial Characterization<br />

The chemical, physical and mechanical<br />

characterization of both types of materials<br />

was carried out using the following<br />

techniques:<br />

1. Chemical composition: Infrared<br />

spectra of the materials were studied<br />

in the 450 to 4000 cm −1 region,<br />

according to ASTM E 1252 standard.<br />

6 For these measurements,<br />

powder samples scraped from the<br />

surfaces of test specimens were<br />

mixed with dry spectroscopic-grade<br />

potassium bromide and pressed into<br />

pellets. 32 scans were collected and<br />

averaged at a spectral resolution of<br />

4 cm −1 , in a Thermo Scientific Nicolet<br />

spectroscope. The glass fibre content<br />

was determined by the calcination<br />

method described in ASTM D 3171<br />

standard. 7<br />

2. Physical properties: Density was<br />

measured according to ISO 1183<br />

standard 8 (immersion method).<br />

Glass transition temperature (T g )<br />

was determined by DMA, in accordance<br />

with ISO 6721 standard. 9<br />

Three-point bending type clamped<br />

specimens of 5 × 15 × 60 mm were<br />

tested at constant frequency of 1 Hz<br />

and strain amplitude of 15 μm, using<br />

a DMA analyser. The analysis was<br />

carried out from room temperature<br />

up to 200°C, at a rate of 2°C/min.<br />

Three replicates were tested for<br />

each type of material.<br />

3. Mechanical properties: Tensile tests<br />

were conducted according to ISO<br />

527—parts 1 and 5 standard 10 in<br />

rectangular test specimens (5 × 25<br />

× 300 mm), without end tabs, using<br />

an universal testing machine with<br />

a load capacity of 100 kN. Threepoint<br />

bending flexural tests were<br />

performed according to ISO 14125<br />

standard, 11 in rectangular test specimens<br />

(5 × 15 × 150 mm) with a span<br />

of 100 mm using a system constituted<br />

by a hydraulic press with a 10<br />

kN load capacity. Interlaminar shear<br />

tests were carried out in accordance<br />

with ASTM D 2344 standard 12 in<br />

rectangular test specimens (5 × 10 ×<br />

30 mm), loaded in a 20 mm span with<br />

the same system used in the bending<br />

tests. Compressive properties were<br />

determined according to ASTM D<br />

695 standard 13 in rectangular specimens<br />

(5 × 10 × 30 mm).<br />

Exposure Environments<br />

In order to study the potential degradation<br />

of the two types of profiles in<br />

typical environments of civil engineering<br />

applications, test specimens were<br />

subjected to the exposure conditions<br />

described in Table 1.<br />

The experimental procedures used<br />

in the immersion ageing conditions,<br />

both in water and in saltwater, were<br />

based on ISO 175 standard, 14 with the<br />

concentration of salt in the saltwater<br />

medium being in agreement with<br />

ASTM D 1141 standard 15 —for both<br />

media, the cut edges of the test specimens<br />

were completely immersed. The<br />

ageing performed in the continuous<br />

condensation chamber was carried out<br />

according to the procedures described<br />

in ISO 6270 standard. 16<br />

1 mm<br />

Mats Rovings Mats Rovings<br />

VE profile<br />

UP profile<br />

Fig. 1: Cross section and fibres architecture of UP and VE profiles (UP = Unsaturated<br />

polyester resin; VE = Vinylester resin)<br />

Fig. 2: Outer mats and inner rovings of a burnt laminate (VE profile)<br />

Experimental Characterization after<br />

Hygrothermal Ageing<br />

After exposure to the different ageing<br />

conditions described in Table 1, batches<br />

of aged test specimens obtained from<br />

each type of profile were subjected<br />

to the following characterization<br />

techniques:<br />

1. Mass changes: Control specimens<br />

with geometry similar to that of specimens<br />

used in DMA were removed<br />

periodically from the different<br />

exposure environments in order to<br />

evaluate their mass changes. After<br />

removal from the exposure environments,<br />

the surface of the specimens<br />

was dried with a cloth in order to<br />

remove any residual free moisture.<br />

Specimens were then immediately<br />

weighed using a 0,0001 g precision<br />

scale.<br />

2. Dynamic mechanical analysis: The<br />

T g was measured according to the<br />

same procedure used in the initial<br />

characterization tests. Three replicates<br />

were tested for each type of<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 371


material and ageing condition (duration<br />

and exposure environment).<br />

3. Mechanical behaviour: Tensile, flexural<br />

and interlaminar shear tests<br />

were performed according to the<br />

above-mentioned standards. At least<br />

five replicates were tested in the longitudinal<br />

direction for each material<br />

and ageing condition.<br />

Excluding the study of the mass<br />

changes, after being removed from<br />

the different exposure environments<br />

and prior to further testing, specimens<br />

were placed inside polyethylene bags.<br />

These were hermetically closed, in<br />

order to maintain the moisture content<br />

of the material, and then placed inside<br />

a room with temperature controlled at<br />

20 (±2)°C. Prior to testing, specimens<br />

were removed from the polyethylene<br />

bags and immediately tested without<br />

any further conditioning.<br />

Results and Discussion<br />

Initial Characterization<br />

The results of initial chemical, physical<br />

and mechanical characterization of<br />

both profiles are listed in Table 2.<br />

Chemical composition determined by<br />

Fourier transform infrared spectroscopy<br />

(FTIR) (Fig. 3) showed little differences<br />

between the two materials<br />

analysed. In fact, the spectra of both<br />

profiles were quite similar. The localizations<br />

and intensity of the peaks confirmed<br />

the presence of ester, as well<br />

as aromatic and aliphatic structures,<br />

which are common in the molecular<br />

structure of both unsaturated polyester<br />

and vinylester resins. FTIR spectra<br />

also showed the existence of calcium<br />

carbonate (as filler) and silica (from<br />

the glass fibres).<br />

The glass fibre content and density<br />

of the VE profile were slightly higher<br />

than those of the UP profile. On the<br />

other hand, the T g (obtained from both<br />

the storage modulus, E′, and the loss<br />

factor, tand) of the VE profile was<br />

lower than that of the UP profile. From<br />

the mechanical aspect point of view,<br />

the UP and VE profiles started to lose<br />

their initial performance at temperatures<br />

above 108 and 99°C, respectively<br />

(Fig. 4).<br />

With regard to the mechanical behaviour,<br />

in all characterization tests (tension,<br />

flexure, interlaminar shear and<br />

compression), both types of profiles<br />

exhibited a well defined and typical<br />

linear elastic behaviour up to failure. A<br />

comparative analysis of the mechanical<br />

behaviour exhibited by both profiles<br />

shows that they are quite similar in<br />

their tensile properties (both strength,<br />

s tu,x , and modulus, E t,x ) and interlaminar<br />

shear strength (s u,sbs ). However,<br />

in flexure, the VE profile showed<br />

superior initial performance, for both<br />

strength (s fu,x ,) and stiffness (E f,x ).<br />

Owing to the relatively high scatter in<br />

the results obtained for the compressive<br />

strength (s cu,x ), it was decided to<br />

Type of exposure Duration Conditions<br />

Immersion in water<br />

(W-20), (W-40), (W-60)<br />

Immersion in saltwater<br />

(S-20), (S-40), (S-60)<br />

Continuous condensation<br />

(CC-40)<br />

Table 1: Exposure ageing conditions<br />

Absorbance Absorbance<br />

1,0<br />

0,8<br />

0,6<br />

0,4<br />

0,2<br />

0,0<br />

1,0<br />

0,8<br />

0,6<br />

0,4<br />

0,2<br />

0,0<br />

UP profile<br />

VE profile<br />

3, 6, 9 and 12 months<br />

3, 6 and 9 months<br />

Fig. 3: FTIR spectra of UP and VE profiles<br />

3000 2000 1000<br />

Wavenumbers (cm –1 )<br />

Composition: demineralized water<br />

Temperatures: 20 (±2)°C,<br />

40 (±1)°C and 60 (±1)°C<br />

Composition: 35 g/L NaCl<br />

Temperatures: 20 (±2)°C,<br />

40 (±1)°C and 60 (±1)°C<br />

Temperature: 40 (±2)°C<br />

Relative humidity: 100%<br />

Property Test method Profile UP Profile VE<br />

Chemical<br />

composition<br />

FTIR<br />

FTIR spectra consistent with unsaturated<br />

polyester or vinylester, with presence<br />

of calcium carbonate and silica<br />

Glass fibre<br />

content (%)<br />

Calcination 68,4 ± 1,8 68,7 ± 0,4<br />

Density (g/cm 3 ) Immersion 1,869 ± 0,113 2,028 ± 0,052<br />

T g (°C) DMA E′ initial 107,9 ± 10,8 98,6 ± 7<br />

tand 146 ± 2,3 126,9 ± 2,3<br />

Mechanical<br />

properties<br />

Tension s tu,x (MPa) 406 ± 31 393 ± 51<br />

E t,x (GPa) 37,6 ± 2,6 38,9 ± 4,1<br />

Flexure s fu,x (MPa) 417 ± 65 537 ± 73<br />

E f,x (GPa) 20,0 ± 6,9 28,4 ± 3,4<br />

Interlaminar<br />

shear<br />

s u,sbs (MPa) 38,5 ± 2,7 39,2 ± 4,2<br />

Compression s cu,x (MPa) 280 ± 123 360 ± 131<br />

Table 2: Initial physical, chemical and mechanical properties<br />

exclude the compressive properties<br />

from this durability study.<br />

Moisture Uptake with Hygrothermal<br />

Ageing<br />

Figure 5 illustrates the mass variation<br />

exhibited by both profiles for<br />

the different immersion media (solutions<br />

of demineralized water, W,<br />

and saltwater, S) and temperatures<br />

372 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


E′(GPa)<br />

(a)<br />

E′(GPa)<br />

(b)<br />

Mass variation (%)<br />

1,50<br />

1,25<br />

1,00<br />

0,75<br />

0,50<br />

0,25<br />

0,00<br />

0<br />

(a)<br />

Mass variation (%)<br />

(b)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

25<br />

Un-aged<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1,50<br />

1,25<br />

1,00<br />

0,75<br />

0,50<br />

0,25<br />

0,00<br />

UP profile<br />

VE profile<br />

0<br />

50 75<br />

UP profile<br />

1000 2000 3000 4000 5000 6000 7000 8000 9000<br />

VE profile<br />

0<br />

100 125 150 175 200<br />

Temperature (°C)<br />

Temperature (°C)<br />

Exposure period (h)<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

1000 2000 3000 4000 5000 6000 7000 8000 9000<br />

Exposure period (h)<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

Fig. 5: Mass variation of UP (a) and VE (b) profiles for different hygrothermal ageing<br />

conditions<br />

0,2<br />

0,16<br />

0,12<br />

0,08<br />

0,04<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

T g,onset<br />

0<br />

25 50 75 100 125 150 175 200<br />

Un-aged W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

Fig. 4: DMA 3-point bending curves of UP (a) and VE (b) profiles before and after<br />

hygrothermal ageing<br />

0,2<br />

0,16<br />

0,12<br />

0,08<br />

0,04<br />

Tan d (–)<br />

Tan d (–)<br />

(20, 40 and 60°C), as well as in continuous<br />

condensation (CC) at 40°C.<br />

Figure 5 shows that the evolution of all<br />

mass variation curves follow roughly<br />

a Fickian response (i.e. a fast initial<br />

mass gain that slows as saturation<br />

approaches), with rates of mass uptake<br />

increasing with temperature, particularly<br />

in the beginning of the exposure. It<br />

can also be seen that, for similar ageing<br />

conditions (immersion media and temperatures),<br />

the comparison of the mass<br />

variation exhibited by both profiles<br />

depicts significant differences, with the<br />

mass uptake for the VE profile being<br />

considerably lower than that exhibited<br />

by the UP profile for all hygrothermal<br />

ageing conditions; these differences,<br />

already reported by Chin et al. 17 , stem<br />

mainly from the distinct water absorption<br />

capacities of both resin systems, in<br />

particular, the higher hydrolytic stability<br />

of VE resins. Figure 5 also shows<br />

that, for similar temperatures, mass<br />

uptake in saltwater was always lower<br />

than that in demineralized water, for<br />

both UP and VE profiles. Finally, one<br />

can readily observe that the increasing<br />

immersion temperature does not<br />

have a direct correlation with the<br />

increased level of mass uptake and<br />

this result should be attributed to the<br />

potential mass loss by extraction of<br />

low molecular components, an effect<br />

that is expected to increase with the<br />

immersion temperature. In fact, weight<br />

changes in these ageing processes usually<br />

result from a balance between the<br />

water uptake due to moisture ingress<br />

and the loss of material. When compared<br />

with the immersion in demineralized<br />

water at 40°C, under continuous<br />

condensation at 40°C both materials<br />

exhibited a higher initial weight gain,<br />

although for longer periods, weight<br />

gains for those environments became<br />

quite similar.<br />

DMA after Hygrothermal Ageing<br />

Figure 4 shows the results of DMA<br />

after 12 months of immersion and 9<br />

months of continuous condensation—<br />

for each condition, only one curve<br />

corresponding to a representative<br />

specimen is plotted. The left axis represents<br />

the variation of E′ curves with<br />

temperature, which exhibit a characteristic<br />

“step” in the glass transition<br />

region; the right axis shows the corresponding<br />

tand curves, which present a<br />

typical peak in that region.<br />

The variation in the behaviour exhibited<br />

by the E′ curves at the transition<br />

region reflects mainly the changes<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 373


in the polymer matrix performance,<br />

which progresses from a glassy state<br />

to an elastomeric state, characteristic<br />

of its viscoelastic nature. In fact, the<br />

reinforced material (in this case, the<br />

glass fibres) does not suffer a stiffness<br />

reduction in this temperature range.<br />

Therefore, the DMA technique indicates<br />

the contribution of the viscoelastic<br />

nature of the matrix for the overall<br />

behaviour of the composite, and in the<br />

present study, this information is useful<br />

to help understand the actual influence<br />

of the matrix nature on the durability<br />

of the composite. In addition, the quality<br />

of the fibre–matrix interface can<br />

also influence DMA results.<br />

Figure 6 plots, in summary, the variation<br />

of the T g of both profiles (mean<br />

value ± standard deviation, determined<br />

based on the onset of E′), as a function<br />

of the type and duration of exposure.<br />

For the UP profile, immersion in<br />

both demineralized water and saltwater<br />

caused, in general, a decrease<br />

in the value of T g . It is seen that the<br />

maximum reduction in T g occurred for<br />

specimens immersed in demineralized<br />

water, whereas the minimum reduction<br />

corresponds to immersion in saltwater.<br />

After 12 months of exposure,<br />

T g seems to increase with the immersion<br />

temperature for both media; for<br />

saltwater immersion at 60°C, the T g<br />

becomes even higher than that of the<br />

un-aged material, most likely due to<br />

a post-curing phenomenon, induced<br />

Tg (E') [°C]<br />

(a)<br />

Tg (E') [°C]<br />

(b)<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

UP profile<br />

VE profile<br />

by the increased temperature. For the<br />

UP profile, the tand curves for continuous<br />

condensation, and immersion<br />

in demineralized water at 60°C show<br />

the appearance of a second “peak” at<br />

higher temperatures (Fig. 4). The occurrence<br />

of this second peak, associated to<br />

the widening of its base, suggests that<br />

the ageing of the material involves a<br />

plasticization mechanism. The occurrence<br />

of two “peaks” in the tanδ curve<br />

may be attributed to the different<br />

mobility of two kinds of segments in<br />

the polymeric matrix, caused by their<br />

different extents of plasticization.<br />

For the VE profile, the variation of T g<br />

was less dependent on the immersion<br />

temperature and, in general, its variation<br />

was less significant than that verified<br />

in the UP profile. The tand curves<br />

for the VE profile did not show any<br />

widening, suggesting that the molecular<br />

structure did not suffer significant<br />

changes. The only exceptions were the<br />

immersions at 60°C in both media, in<br />

which an asymmetry could be observed<br />

in the configuration of the tand curves,<br />

near their maximum value. This result is<br />

consistent with the lower water uptake<br />

ability exhibited by this material, when<br />

compared with the UP profile.<br />

Water uptake by unsaturated polyester<br />

and vinylester composites is known to<br />

cause plasticization in the short term<br />

and hydrolysis over the long term<br />

through attack of the ester linkages. 18<br />

As the ester group is located in the<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

Initial 3 months 6 months 9 months 12 months<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

Initial 3 months 6 months 9 months 12 months<br />

Fig. 6: Glass transition temperature of UP (a, c) and VE (b, d) profiles after<br />

hygrothermal ageing<br />

middle of the molecular structure of<br />

polyester, and in the ends of the molecular<br />

structure of vinylester, in principle,<br />

the later resin is more resistant to the<br />

above-mentioned plasticization mechanisms.<br />

Both these phenomena induce<br />

higher levels of molecular mobility,<br />

resulting in a consequent decrease in<br />

the T g , although such decrease can<br />

often be offset through residual curing<br />

of the resins in aqueous media. These<br />

competing phenomena result in fluctuations<br />

in the T g as a function of the<br />

exposure period; in the experiments<br />

reported herein, such behaviour was<br />

shown, in particular, by the UP profile.<br />

Mechanical Performance after<br />

Hygrothermal Ageing<br />

Tensile Properties<br />

The results obtained from tensile tests<br />

on both materials, namely, the tensile<br />

strength and the tensile modulus as<br />

a function of time and hygrothermal<br />

conditions, are presented in Fig. 7.<br />

Figure 7 shows that for all ageing conditions<br />

and for both materials there<br />

was an overall decrease in the average<br />

tensile strength with the duration of<br />

exposure (the only exception was the<br />

VE profile after 12 months of exposure<br />

in demineralized water at 20°C).<br />

As expected, the level of degradation<br />

of the tensile strength increased consistently<br />

with the temperature of the<br />

immersion medium, with maximum<br />

reductions occurring at 60°C—after 12<br />

months of exposure, the lowest levels<br />

of retention were 64 and 75% for the<br />

UP and VE profiles, both immersed<br />

in demineralized water. The general<br />

higher aggressiveness of demineralized<br />

water compared to saltwater was<br />

consistent with previous investigations<br />

(e.g. those reported by Van de Velde<br />

and Kiekens 19 ) and could be attributed<br />

to osmotic effects. For all hygrothermal<br />

ageing conditions and periods of exposure,<br />

the tensile strength retention of<br />

the VE profile was consistently higher<br />

than that of the UP profile. Chu et al. 20<br />

reported tensile strength reductions<br />

in pultruded E-glass vinylester laminates<br />

that follow the overall trend of<br />

the present tests—the strength retention<br />

presented by those authors was<br />

considerably smaller, but so was the<br />

thickness of the tested material and,<br />

consequently, the maximum moisture<br />

uptake.<br />

The variation exhibited by the average<br />

tensile modulus with the exposure<br />

period (Fig. 7), which was much<br />

more irregular and associated with<br />

374 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Tensile strength (MPa)<br />

(a)<br />

Tensile modulus (GPa)<br />

(c)<br />

500<br />

UP profile<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

W-20 W-40 W-60 S-20<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

S-40 S-60 CC-40<br />

Initial 3 months 6 months 9 months 12 months<br />

Tensile modulus (GPa) Tensile strength (MPa)<br />

(b)<br />

500<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

VE profile<br />

W-20 W-40 W-60 S-20<br />

S-40 S-60 CC-40<br />

Initial 3 months 6 months 9 months 12 months<br />

UP profile<br />

55<br />

50<br />

VE profile<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40 W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

Initial 3 months 6 months 9 months 12 months (d) Initial 3 months 6 months 9 months 12 months<br />

Fig. 7- Tensile strength and modulus of UP (a, c) and VE (b, d) profiles after hygrothermal ageing<br />

higher coefficients of variation (when<br />

compared to the tensile strength),<br />

makes it more difficult to establish<br />

systematic analyses and comparisons<br />

between the two materials. However,<br />

for all exposure conditions and durations,<br />

one can conclude that the stiffness<br />

retention was considerably higher<br />

than the corresponding strength retention—the<br />

minimum levels of retention<br />

after 12 months were 76% for the UP<br />

profile (saltwater at 20°C) and 85%<br />

for the VE profile (saltwater at 60°C).<br />

In addition, and similar to the tensile<br />

strength, the stiffness retention of the<br />

VE profile was always higher than that<br />

exhibited by the UP profile. Finally,<br />

when compared to strength, stiffness<br />

retention appeared to be much more<br />

insensitive to the temperature and<br />

composition of the immersion media.<br />

For both profiles the strength and stiffness<br />

retention of specimens immersed<br />

in demineralized water at 40°C was<br />

comparable to that of specimens under<br />

continuous condensation (particularly<br />

for the VE profile) and this result<br />

agrees well with the water uptake<br />

measurements.<br />

Flexural Properties<br />

The results obtained from flexural<br />

tests, namely the flexural strength and<br />

the flexural modulus, on both materials<br />

are presented as a function of time<br />

and hygrothermal conditions in Fig. 8.<br />

In general, the level of degradation of<br />

the flexural strength of both profiles<br />

increased with the temperature of the<br />

immersion medium, being maximum<br />

at 60°C—this result was in agreement<br />

with the behaviour already reported<br />

for the tensile strength and with results<br />

obtained by other authors with GFRP<br />

pultruded profiles. 19,21 As for the tensile<br />

strength, for similar temperatures,<br />

strength reduction in demineralized<br />

water was usually lower than that in<br />

saltwater (it should be mentioned that<br />

this latter result differs from those<br />

reported by Liao et al. 22 ) Unlike the<br />

behaviour exhibited in the tensile tests<br />

(essentially dominated by the fibres),<br />

in flexure (also significantly influenced<br />

by the matrix and the fibre–matrix<br />

interface) the VE profile did not present<br />

a better mechanical performance<br />

than the UP profile for all ageing<br />

conditions; in fact, for most exposure<br />

conditions and durations, the strength<br />

retention of the VE profile was considerably<br />

smaller than that of the UP profile—after<br />

12 months of exposure, the<br />

lowest levels of retention were 66%<br />

for the UP profile and 58% for the<br />

VE profile, both immersed in demineralized<br />

water. This result may be due,<br />

at least to some extent, to some postcuring<br />

effect on the UP profile, which<br />

was identified in the DMA tests, and, in<br />

addition, to the considerable scatter of<br />

the results obtained in the initial characterization<br />

flexural tests. In this regard,<br />

it is worth mentioning that Kootsookos<br />

and Mouritz 23 also reported higher<br />

flexural strength retention in moulded<br />

glass–polyester laminates, compared<br />

with glass–vinylester laminates having<br />

similar fibre architectures.<br />

Flexural modulus after 12 months of<br />

exposure decreased in all ageing conditions,<br />

for both profiles (Fig. 8). Similar<br />

to strength, in general, stiffness retention<br />

in demineralized water was lower<br />

than that in saltwater and, in addition,<br />

the UP profile presented better<br />

performance than the VE profile—the<br />

lowest levels of retention were 69%<br />

for the UP profile and 58% for the VE<br />

profile, both immersed in demineralized<br />

water. Similar to the tensile modulus,<br />

the flexural stiffness retention<br />

appeared to be more insensitive to the<br />

temperature of the immersion media,<br />

when compared to strength.<br />

As for tensile performance, the variation<br />

of flexural properties in specimens<br />

subjected to continuous condensation<br />

was roughly similar to that of<br />

specimens immersed in demineralized<br />

water at 40°C.<br />

Interlaminar Shear Strength<br />

Figure 9 illustrates the variation of the<br />

interlaminar shear strength (a matrix<br />

dominated property) of both profiles<br />

as a function of the hygrothermal conditions<br />

and the period of exposure.<br />

Figure 9 shows that very significant<br />

reductions occurred in the interlaminar<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 375


600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

(a)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Flexural strength (MPa)<br />

Flexural modulus (GPa)<br />

UP profile<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

700<br />

VE profile<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

Initial 3 months 6 months 9 months 12 months (b) Initial 3 months 6 months 9 months 12 months<br />

UP profile<br />

0<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

W-20 W-40 W-60 S-20 S-40 S-60 CC-40<br />

(c)<br />

Initial 3 months 6 months 9 months 12 months (d) Initial 3 months 6 months 9 months 12 months<br />

Fig. 8– Flexural strength and modulus of UP (a, c) and VE (b, d) profiles after hygrothermal ageing<br />

Flexural modulus (GPa) Flexural strength (MPa)<br />

35<br />

VE profile<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

shear strength of both profiles as a<br />

function of time. As for the other two<br />

mechanical tests, and similar to results<br />

reported earlier by Kharbari, 24 the<br />

interlaminar shear strength retention<br />

decreased consistently with the immersion<br />

temperature. After 12 months<br />

of exposure at 60°C, the strength<br />

retention in demineralized water was<br />

approximately 55% for the UP profile<br />

and 61% for the VE profile. For<br />

saltwater immersion, the corresponding<br />

values were slightly higher, with<br />

strength retentions of 60 and 62% for<br />

the UP and VE profiles, respectively.<br />

As for the two other mechanical tests,<br />

the variation of the interlaminar shear<br />

strength for immersion in demineralized<br />

water at 40°C was roughly<br />

analogous to that under continuous<br />

condensation. Strength reductions<br />

exhibited by the VE profile immersed<br />

in demineralized water at the three<br />

different temperatures followed a<br />

trend similar to those reported by Chu<br />

et al. 20 —these authors obtained higher<br />

strength retentions but, as already discussed,<br />

the results of the two studies<br />

are not directly comparable. Finally,<br />

the better performance exhibited by<br />

the VE profile for all hygrothermal<br />

ageing conditions and periods of exposure<br />

is outlined.<br />

Conclusion<br />

This paper presented results of an<br />

ongoing research project on the<br />

environmental degradation suffered<br />

by GFRP pultruded profiles made of<br />

either UP or VE resins, with similar<br />

fibre contents and architectures. On<br />

the basis of results obtained for an<br />

exposure of 12 months in demineralized<br />

water and saltwater at 20, 40 and<br />

60°C, as well as 9 months in continuous<br />

condensation at 40°C, the following<br />

conclusions can be arrived at:<br />

Interl. shear strength (MPa) Interl. shear strength (MPa)<br />

(a)<br />

50<br />

UP profile<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

W-20 W-40<br />

(b)<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Initial 3 months 6 months 9 months 12 months<br />

VE profile<br />

W-20<br />

W-40<br />

W-60<br />

W-60<br />

1. The water uptake capacity of GFRP<br />

profiles and their temperature<br />

dependency are strongly dependent<br />

on the nature of the polymeric<br />

matrix—for similar ageing conditions,<br />

the VE profile exhibited considerably<br />

lower mass uptake than<br />

the UP profile.<br />

2.The UP profile was the one that<br />

presented signs of plasticization in<br />

S-20<br />

S-20<br />

S-40<br />

S-40<br />

S-60<br />

S-60<br />

CC-40<br />

CC-40<br />

Initial 3 months 6 months 9 months 12 months<br />

Fig. 9: Interlaminar shear strength of UP (a) and VE (b) profiles after hygrothermal<br />

ageing<br />

376 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


DMA, and as a consequence of such<br />

plasticization mechanism the values<br />

of T g in aged specimens suffered a<br />

general reduction. In spite of this,<br />

for some ageing conditions, namely<br />

for higher temperatures, such reduction<br />

was offset because of the occurrence<br />

of resin post-curing. These<br />

competing phenomena resulted in<br />

fluctuations in the T g as a function<br />

of the period of exposure. For the<br />

VE profile, variations of T g were<br />

much lower, with DMA not suggesting<br />

any appreciable changes in the<br />

molecular structure.<br />

3. The mechanical properties of GFRP<br />

profiles constituted by both types<br />

of resins were noticeably affected,<br />

even for immersion at 20°C (Fig. 10).<br />

For most of the hygrothermal ageing<br />

conditions and periods of exposure,<br />

the retention of both tensile strength<br />

(a fibre-dominated mechanical<br />

property) and interlaminar shear<br />

strength (matrix dominated) of the<br />

VE profile was considerably higher<br />

than that of the UP profile—degradation<br />

increased with the temperature<br />

of the immersion medium, with<br />

demineralized water being generally<br />

more aggressive than saltwater. In<br />

flexure, the tendency of the results<br />

was not so clear and, to some extent<br />

was contradictory to results of the<br />

other mechanical tests, as the VE<br />

profile showed a generally worse<br />

performance than the UP profile. It<br />

is believed that flexural results may<br />

have been influenced by the postcuring<br />

phenomen on observed in the<br />

UP profile.<br />

4. The above-mentioned degradation<br />

was mainly due to physical phenomena,<br />

such as plasticization of the<br />

polymeric matrix, since no appreciable<br />

chemical degradation was<br />

detected through FTIR analyses. 25,26<br />

Nevertheless, this degradation may<br />

influence the use of GFRP profiles<br />

in wet environments (structures<br />

placed underwater or subjected to<br />

Property retention (%)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Tensile<br />

strength<br />

Tensile<br />

modulus<br />

high levels of moisture) and especially<br />

in tropical zones (where, in<br />

addition, temperatures are high),<br />

where the use of different resin<br />

systems and/or superficial protections<br />

(such as paintings or gel coats)<br />

shall be considered for improved<br />

performance.<br />

The tendencies stated in the aforementioned<br />

conclusions will be assessed<br />

and eventually confirmed in the forthcoming<br />

experiments to be carried out<br />

within this research project. Additional<br />

experiments are also being carried out<br />

in order to evaluate the reversibility of<br />

the degradation suffered by both profiles—in<br />

these new experiments, specimens<br />

will be submitted to mechanical<br />

tests after being dried, as against the<br />

present experiments, in which they<br />

were tested in a saturated state. The<br />

next steps will also include the development<br />

of analytical models in order<br />

to simulate the observed degradation<br />

suffered by both types of profiles.<br />

Acknowledgements<br />

The authors wish to acknowledge the support<br />

of FCT, ICIST and Agência da Inovação<br />

(Grant No. 2009/003456) for funding the<br />

research and also STEP and ALTO for supplying<br />

the GFRP profiles used in the experimental<br />

investigations.<br />

References<br />

[1] Correia JR. GFRP pultruded profiles in<br />

civil engineering: hybrid solutions, bonded connections<br />

and fire behaviour, PhD Thesis, IST,<br />

Technical University of Lisbon, 2008.<br />

[2] Correia JR, Cabral-Fonseca S, Branco FA,<br />

Ferreira J, Eusébio MI, Rodrigues MP. Durability<br />

of glass fibre reinforced polyester (GFRP) pultruded<br />

profiles for construction applications.<br />

Mech. Compos. Mater. 2006; 42(4): 325–338.<br />

[3] Keller T, Bai Y, Vallée T. Long-term performance<br />

of a glass-fiber-reinforced polymer truss<br />

bridge. J. Compos. Const. 2007; 11(1): 99–108.<br />

[4] Kharbari VM, Chin JW, Hunston D,<br />

Benmokrane B, Juska T, Morgan R, Lesko JJ,<br />

Flexural<br />

strength<br />

Flexural<br />

modulus<br />

UP-W-20 VE-W20 UP-S-20 VE-S-20<br />

Int. shear<br />

strength<br />

Fig. 10: Retention of mechanical properties after 12 months of immersion at 20°C<br />

Sorathia U, Reynaud D. Durability gap analysis<br />

for fiber-reinforced polymer composites in<br />

civil infrastructure. J. Compos. Const. 2003; 7:<br />

238–247.<br />

[5] Harries KA, Porter MA, Busel JP. FRP materials<br />

and concrete research needs. Concr. Int.,<br />

ACI 2003; October: 69–74.<br />

[6] ASTM~E~1252. Standard Practice for General<br />

Techniques for Obtaining Spectra Qualitative<br />

Analysis. American Society for Testing and<br />

Materials, West Conshohocken, PA, 1998.<br />

[7] ASTM~D~3171. Standard Test Methods for<br />

Constituent Content of Composite Materials.<br />

American Society for Testing and Materials,<br />

West Conshohocken, PA, 2009.<br />

[8] ISO~1183—Plastics. Methods for<br />

Determining the Density of Non-Cellular Plastics.<br />

Part 1: Immersion Method, Liquid Pyknometer<br />

Method And Titration Method. International<br />

Organization for Standardization, Brussels, 2004.<br />

[9] ISO~6721—Plastics. Determination of Dynamic<br />

Mechanical Properties. Part 1: General Principles;<br />

Part 5: Flexural Vibration- Non-Resonance<br />

Method. International Organization for<br />

Standardization, Brussels, 2001.<br />

[10] ISO~527—Plastics. Determination of Tensile<br />

Properties. Part 1: General Principles; Part~5: Test<br />

Conditions for Unidirectional Fiber-Reinforced<br />

Plastic Composites, International Standards<br />

Organization, Brussels, 1997.<br />

[11] ISO~14125. Fiber-Reinforced Plastic Composites—Determination<br />

of Flexural Properties,<br />

International Standards Organization,<br />

Brussels, 1998.<br />

[12] ASTM~D~2344. Standard Test for Short-<br />

Beam Strength of Polymer Matrix Composite<br />

Materials and Their Laminates, American Society<br />

for Testing and Materials, West Conshohocken,<br />

PA, 2006.<br />

[13] ASTM~D~695. Standard Test Method for<br />

Compressive Properties of Rigid Plastics.<br />

American Society for Testing and Materials,<br />

West Conshohocken, PA, 2008.<br />

[14] ISO~175—Plastics. Methods of Test for the<br />

Determination of the Effects of Immersion in<br />

Liquid Chemicals. International Organisation<br />

for Standardization, Brussels, 1999.<br />

[15] ASTM~D~1141. Standard Practice for<br />

the Preparation of Substitute Ocean Water.<br />

American Society for Testing and Materials,<br />

West Conshohocken, PA, 2008.<br />

[16] ISO~6270—Paints and varnishes. Determination<br />

of Resistance to Humidity. Part 1:<br />

Continuous Condensation, International<br />

Standards Organization, Brussels, 1998.<br />

[17] Chin JW, Nguyen T, Aouadi K. Effects of<br />

environmental exposure on fiber- reinforced<br />

plastic (FRP) materials used in construction.<br />

J. Compos. Technol. Res. 1997; 19(4):<br />

205–213.<br />

[18] Chu W, Karbhari VM. Effect of water<br />

sorption on performance of pultruded E-glass/<br />

vinylester composites. J. Mater. Civil Eng. 2005;<br />

17(1): 63–71.<br />

[19] Van de Velde K, Kiekens P. Effects of chemical<br />

environments on pultruded E-glass reinforced<br />

polyesters. J. Compos. Technol. Res. 2001;<br />

23(2): 92–101.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 377


[20] Chu W, Wu L, Kharbari VM. Durability<br />

evaluation of moderate temperature cured<br />

E-glass/vinlyester systems. Compos. Struct. 2004;<br />

66(1–4): 367–376.<br />

[21] Nishizaki I, Meiarashi S. Long-term deterioration<br />

of GFRP in water and moist environment.<br />

J. Compos. Const. 2002; 6(1): 21–27.<br />

[22] Liao K, Schultheisz CR, Hunston DL.<br />

Effects of environmental aging on the properties<br />

of pultruded GFRP. Compos. B: Eng. 1999; 30(5):<br />

485–493.<br />

[23] Kootsookos A, Mouritz AP. Seawater durability<br />

of glass- and carbon-polymer composites.<br />

Compos. Sci. Technol. 2004; 64(10–11): 1503–1511.<br />

[24] Kharbari VM. E-Glass/vinylester composites<br />

in aqueous environments: effects on shortbeam<br />

shear strength. J. Compos. Const. 2004;<br />

8(2): 148–156.<br />

[25] Costa R. Durability of Glass Fibre<br />

Reinforced Polyester Pultruded Profiles. MSc<br />

Thesis, IST, Technical University of Lisbon, 2009<br />

(in Portuguese).<br />

[26] Carreiro A. Durability of Glass Fibre<br />

Reinforced Vinylester Pultruded Profiles. MSc<br />

Thesis, IST, Technical University of Lisbon, <strong>2010</strong><br />

(in Portuguese).<br />

International Association for Bridge<br />

and Structural Engineering<br />

IABSE Conference Cairo Egypt<br />

May 7-9, 2012<br />

Global Thinking in Structural Engineering:<br />

Recent Achievements<br />

Organised by<br />

The Egyptian Group of IABSE<br />

The Egyptian Society of Engineers<br />

More information: secretariat@iabse.org<br />

378 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Evaluation of a Life Prediction Model and Environmental<br />

Effects of Fatigue for Glass Fiber Composite Materials<br />

David Brian Dittenber, Ph.D. Student; GangaRao V.S. Hota, Prof., Director of Constructed Facilities Center; Civil Eng. Dept.,<br />

West Virginia University, USA. Contact: ddittenb@mix.wvu.edu<br />

Abstract<br />

Fiber-reinforced polymer (FRP) composites are being utilized in an increasing<br />

number of applications in structural industries. Establishing values for longterm<br />

fatigue response is essential for widespread use of FRP composites. The<br />

variety of available fibers, matrix materials, and manufacturing processes makes<br />

the fatigue response difficult to predict without extensive empirical testing.<br />

A proposed fatigue life prediction model uses the internal strain energy release<br />

rate as the metric for predicting fatigue life from a minimum of data points. The<br />

objective of this research was to apply the above model to fatigue data for various<br />

composite coupons and components in order to evaluate its applicability in<br />

predicting fatigue life. The model was found to be able to regularly fit and predict<br />

fatigue data within 5% log error at both coupon and component levels. The<br />

effects of environmental conditions, including 12 MPa pressurized absorption<br />

and fatiguing in salt water and elevated temperatures, were also explored for a<br />

glass/vinyl ester FRP. The results of this research can be used to aid in the design<br />

of numerous structural FRP applications, such as windmill blades, bridge decks,<br />

or deep sea piping.<br />

Keywords: FRP composites; fatigue life prediction; temperature effects; saltwater<br />

environment; component fatigue; strain energy.<br />

Introduction<br />

Fiber-reinforced polymer (FRP)<br />

composites offer an opportunity to<br />

revolutionize and rehabilitate infrastructure<br />

because of their advantages<br />

over traditional structural materials:<br />

higher strength, higher fatigue and<br />

impact resistance, higher corrosive and<br />

environmental resistance, longer service<br />

life, lower installation and maintenance<br />

costs, and more consistent<br />

performance. 1<br />

FRPs are composites that have reinforcement<br />

in the form of fibers and<br />

matrix in the form of a polymeric<br />

material, that is, epoxy, vinyl ester, or<br />

polyester. The most common fiber<br />

reinforcements are glass and carbon<br />

fibers, but the higher cost of carbon<br />

generally prohibits their use for infrastructure.<br />

Glass fiber-reinforced polymers<br />

(GFRPs) are lightweight and<br />

have a good combination of mechanical<br />

performance per unit cost.<br />

Performing extensive experimental<br />

work is one way to fully characterize<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: March 14, <strong>2010</strong><br />

Paper accepted: May 21, <strong>2010</strong><br />

the performance of a material, but<br />

fatigue testing is time consuming and<br />

costly. Therefore, observed behavioral<br />

trends need to be combined with<br />

established physical relationships in<br />

order to produce mathematical models<br />

to predict fatigue life with reasonable<br />

accuracy. Many researchers have developed<br />

fatigue life prediction models for<br />

composites in the past few decades,<br />

but none have been widely adapted in<br />

industry.<br />

Fatigue in Composites<br />

When the weakest location in a composite<br />

fails, the surrounding fiber/<br />

matrix interface experiences increased<br />

local stresses, which can lead to fatigue<br />

damage. The weakest location is typically<br />

caused by material defects, such<br />

as misaligned fibers, voids, or resin-rich<br />

regions. As fatigue damage progresses,<br />

cracks in either the resin or the resin/<br />

matrix interface develop between the<br />

fibers, causing many to become overloaded<br />

and fail. In order to predict<br />

the fatigue behavior of a material, the<br />

remaining strength or the remaining<br />

number of cycles, the material properties<br />

and the environmental/loading<br />

conditions must be related to the various<br />

damage modes that the material is<br />

likely to experience.<br />

When developing a fatigue life prediction<br />

model, it is important to evaluate<br />

how the model is able to handle a variety<br />

of materials and test conditions.<br />

Several researchers have compiled a<br />

large composite material fatigue database<br />

consisting of data from fatigue<br />

testing on wind turbine blade materials.<br />

2,3 Over 190 polymeric composite<br />

materials with over 12 000 individual<br />

coupon tests were conducted. Their<br />

research was focused on materials<br />

with lay-up combinations of 0°, ±45°,<br />

and 0°/±45° fabrics, manufactured by<br />

either hand lay-up or resin transfer<br />

molding (RTM).<br />

Fatigue Model Development<br />

The proposed GFRP fatigue life<br />

model 4 uses the internal strain energy<br />

of the material as the damage metric;<br />

this energy is expended due to damage<br />

in the forms of matrix cracking,<br />

fiber/matrix interface failure, delaminations,<br />

or fiber breakage before rupture.<br />

Strain energy was chosen as the<br />

damage metric because of its ability<br />

to represent multiple damage modes<br />

through one measurement and its high<br />

sensitivity to damage accumulation<br />

due to the squaring of the stress/strain<br />

term. While other researchers 5 have<br />

previously used the strain energy to<br />

predict fatigue life, this more recent<br />

model differs in the following ways:<br />

1. It is laminate-derived instead of<br />

lamina-derived, meaning no ply<br />

mechanics calculations need to be<br />

made.<br />

2. It focuses on utilizing a minimum<br />

number of experimental tests to<br />

determine material coefficients.<br />

3. It relates the strain energy back to<br />

the stress versus number of cycles<br />

curve in order to provide physical<br />

meaning.<br />

4. It is intended for use in industry,<br />

therefore attempts to simplify the<br />

process while still providing a reasonable<br />

fatigue life prediction.<br />

The expenditure of strain energy<br />

occurs in three stages (Fig. 1): Stage I<br />

is a steep curve of energy loss as the<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 379


Expended energy (U)<br />

Failure<br />

Stage I Stage II (Linear) Stage III<br />

N<br />

f<br />

Uo<br />

Uo<br />

2 dU<br />

dN<br />

2 a σ σ<br />

= ( )<br />

=<br />

Model Evaluation<br />

b<br />

( ( max ult ) )<br />

(10)<br />

Fatigue test results were selected from<br />

the composite material fatigue database<br />

2 using the following controlling<br />

criteria:<br />

Cycles to failure (N)<br />

Fig. 1: Three stages of energy expenditure in fatigue<br />

material is initially loaded, and generally<br />

lasts for about 15% of the fatigue<br />

life; Stage II is nearly linear, with the<br />

slope a characteristic of the material<br />

and testing conditions, and generally<br />

lasts for about 75% of the fatigue life;<br />

Stage III is again a steep curve leading<br />

to material rupture over the last 10%<br />

of the fatigue life. 4,6<br />

Equation (1) describes the rate of<br />

the release of strain energy, U j , with<br />

respect to the number of cycles, N j ,<br />

as a function of the mean strain, e m ,<br />

the strain range, e r , the amount of<br />

expended strain energy corresponding<br />

to the current cycle count, and the particular<br />

composite material type, C t . 4 If<br />

the material type is kept constant, then<br />

this relationship supposes that the rate<br />

of release of strain energy with respect<br />

to the number of cycles is a function of<br />

the loading conditions:<br />

dU<br />

j<br />

=<br />

m r t<br />

dN f ( , , U C j, (1)<br />

)<br />

j<br />

The strain energy at any cycle of a<br />

fatigue test can be determined from<br />

the deflection (or strain) and loading<br />

data; for a tension–tension fatigue test,<br />

this is shown in Eq. (2), where P j is<br />

the load and d j is the deflection at any<br />

cycle j:<br />

U<br />

P j<br />

j<br />

j<br />

=<br />

(2)<br />

2<br />

To account for more complex material<br />

properties, it would be acceptable to use<br />

different versions of the strain energy<br />

equation. The strain energy density, W,<br />

could be calculated as shown in Eq. (3)<br />

for elastic plane stress problems. 5<br />

( )<br />

W =<br />

x x<br />

+<br />

y y<br />

+<br />

xy xy<br />

/2 (3)<br />

The strain energy model calculates<br />

the initial strain energy U 0 using the<br />

mean loading stress (s mean ) just prior<br />

to initiating fatigue, as shown in Eq.<br />

(4), where A is the cross sectional area<br />

of the sample, L is the gage length,<br />

and E static is the static modulus of<br />

elasticity. 4<br />

2<br />

AL<br />

U0<br />

= mean<br />

(4)<br />

2Estatic<br />

After analyzing experimental strain<br />

energy release results for several GFRP<br />

fatigue tests, it was observed that the<br />

strain energy expended at ~90% of the<br />

fatigue life (the end of Stage II) was<br />

consistently close to 1,5 times the initial<br />

strain energy, U 0 . The data point at<br />

the end of Stage II can therefore be<br />

defined as (N f , U f ), as shown in Eqs.<br />

(5) and (6). 4<br />

U 15 f<br />

, U<br />

(5)<br />

0<br />

Nf<br />

09 , N<br />

(6)<br />

ult<br />

The energy release rate per cycle is<br />

relatively constant within Stage II and<br />

is dependent on the material type and<br />

loading conditions. Plotting the normalized<br />

strain against the energy release<br />

rate, the data can be curve-fitted with<br />

a power law as shown below, where a<br />

and b are the fatigue coefficients.<br />

b<br />

dU<br />

dN = a ⎛ ε ⎞<br />

⎝ ⎜ ε ⎠<br />

⎟<br />

max (7)<br />

ult<br />

Because of the linearity of the Stage II<br />

energy release, dU/dN can be approximated<br />

as ΔU/ΔN [Eq. (8)]. If we assume<br />

that the strain ratio can be approximated<br />

as equivalent to the stress ratio<br />

due to a linear relationship between<br />

the stress and strain of the material<br />

[Eq. (9)], N f can be calculated as shown<br />

in Eq. (10).<br />

dU<br />

U<br />

15 , U U<br />

= =<br />

N N<br />

N<br />

U<br />

0 0 0<br />

d<br />

f<br />

2Nf<br />

<br />

<br />

max<br />

ult<br />

=<br />

(8)<br />

<br />

max<br />

(9)<br />

<br />

ult<br />

1. Tension–tension testing<br />

2. Stress ratio (s min /s max ), R = 0,1<br />

3. Minimum three load levels (between<br />

25 and 75% of max stress)<br />

4. Minimum five tests run to failure<br />

with at least 100 cycles each<br />

5. E-glass fibers<br />

6. Fatigue rate ≤ 20 Hz<br />

7. Generally ordered results (if s 1 > s 2 ,<br />

then N 1 < N 2 ).<br />

After applying these criteria, test<br />

results for 109 different composites<br />

(1254 individual tests) were analyzed.<br />

Each material listing in the database<br />

included its static modulus of elasticity,<br />

static ultimate strength, and<br />

coupon dimensions, with stress load<br />

levels and number of cycles to failure<br />

for each coupon. Polyester was<br />

the most common matrix material,<br />

but results for vinyl ester, epoxy, and<br />

thermoplastic matrices were included.<br />

Most of the laminates selected were<br />

manufactured by RTM or lay-up techniques<br />

but included a variety of fiber<br />

architectures.<br />

For each composite, the fatigue coefficients<br />

(a and b) were obtained through<br />

a power regression on the criteriaselected<br />

data after plotting the results<br />

of Eqs. (8) and (9) (Fig. 2). Using the<br />

fatigue coefficients, the curve generated<br />

by Eq. (10) was plotted against<br />

the data on a log scale (see Fig. 3).<br />

In order to assess the accuracy of the<br />

model in fitting data, curves of ±5%<br />

log(N f ) and ±10% log(N f ) were generated,<br />

with any data lying within these<br />

envelopes considered to be reasonably<br />

well-modeled (Fig. 3). The number of<br />

data points within each error envelope<br />

was then tallied for each sample.<br />

Results<br />

Composite Material Fatigue<br />

Database<br />

The results of the fatigue database 2<br />

analysis of the model are shown in<br />

Table 1. Since data scatter is as likely<br />

as a poor-fitting model to result in data<br />

380 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Energy release rate<br />

ΔU/ΔN in N-mm/cycle<br />

Normalized max stress (s max /s ult )<br />

10<br />

9<br />

8<br />

ΔU/ΔN<br />

7<br />

Power (ΔU/ΔN)<br />

6<br />

5<br />

y = 2610,44 × 17,67<br />

4<br />

R 2 = 0,989<br />

3<br />

2<br />

1<br />

0<br />

0 0,2 0,4 0,6 0,8<br />

Fig. 2: Typical mate rial energy release rate<br />

0,85<br />

0,75<br />

0,65<br />

0,55<br />

0,45<br />

0,35<br />

Normalized applied stress (s max /s ult )<br />

0,25<br />

0 2 4 6 8 10<br />

Cycles to failure (Log N)<br />

Fig. 3: Typical log error analysis (on glass/vinyl ester composite)<br />

Actual<br />

Model fit<br />

Upper limit (+10%)<br />

Lower limit (−10%)<br />

Upper limit (+5%)<br />

Lower limit (−5%)<br />

Total Within 5% log error Within 10% log error<br />

Number Number Percentage (%) Number Percentage (%)<br />

Full data set 1260 1034 82,1 1222 97<br />

R 2 > 0,9 1201 1003 83,5 1171 97,5<br />

R 2 > 0,95 1099 935 85,1 1078 98,1<br />

R 2 > 0,98 623 565 90,7 621 99,7<br />

Table 1: Log error anal ysis results<br />

points located outside of these error<br />

envelopes, results were also considered<br />

for different correlation coefficient<br />

(R 2 ) values obtained from the initial<br />

power regression. A higher R 2 value<br />

would indicate less scatter; using R 2 ><br />

0,98 (Table 1) the model is able to fit<br />

over 90% of the fatigue data to within<br />

the ±5% error envelope.<br />

Once it had been shown that the<br />

model provided good accuracy at fitting<br />

the data, another analysis was run<br />

to assess its ability to predict coupon<br />

fatigue life. Only composites with polyester<br />

matrices and a total of nine data<br />

points were considered, resulting in 14<br />

laminates and a total of 126 fatigue<br />

tests. The same curve equation and<br />

error envelopes were generated for<br />

each composite using several different<br />

fatigue coefficients. The power regression<br />

was performed using anywhere<br />

from two to nine logically selected<br />

data points and the number of data<br />

points within the error envelopes was<br />

tallied. A curve was also generated<br />

for each composite using only a single<br />

data point (~50% ultimate strength)<br />

and the loose correlation of Eq. (11).<br />

( <br />

)<br />

b = 20 1 fiber volume fraction (11)<br />

Grouping and analyzing the data by one<br />

or more of the following characteristics<br />

at a time produced no noticeable<br />

trends in the values of either the a<br />

or b coefficients: matrix type, common<br />

fiber architecture, manufacturing<br />

process, material thickness, material<br />

cross-sectional area, modulus of elasticity,<br />

testing rate, and ultimate tensile<br />

strength. The differences in matrices<br />

may have had a slight impact on the<br />

b-coefficient (with epoxies averaging<br />

11,8, polyesters 11,9, and vinyl esters<br />

12,1), but was not large enough to be<br />

significant and did not account for the<br />

high degree of variation within each of<br />

those groups.<br />

The linear approximation of Eq. (11)<br />

was obtained by running regressions<br />

on plots of the fiber volume fraction<br />

versus the b-coefficient for several<br />

smaller groupings of RTM polyester-matrix<br />

data (e.g. unidirectional<br />

samples, 0°/45° samples, etc.). Plotting<br />

the b-coefficient against the fiber volume<br />

fraction still does not account for<br />

the high amount of scatter (Fig. 4),<br />

but does allow for a reasonable average<br />

to be obtained. An attempt was<br />

made to normalize the fiber content<br />

with respect to the 0° direction before<br />

plotting against the b-coefficient, but it<br />

did not provide any better an approximation.<br />

Equation (11) provides an<br />

approximation of the b-coefficient for<br />

a majority of samples only if they have<br />

polyester matrices and were manufactured<br />

by RTM; the error can be much<br />

higher for those samples with other<br />

matrices or manufacturing methods.<br />

The percentages of the 126 data points<br />

that fell within the ±5 and ±10% error<br />

envelopes using between one and nine<br />

data points to obtain the fatigue coefficients<br />

are shown in Fig. 5.<br />

Glass/Vinyl Ester Composite<br />

The same logarithmic error envelope<br />

analysis was carried out using fatigue<br />

test results from a glass/vinyl ester composite<br />

material. Six tension– tension<br />

fatigue tests were run at six different<br />

stress levels between 35 and 70% of<br />

the ultimate static strength.<br />

After performing a similar curve fit<br />

analysis as was conducted on the database<br />

materials, it was determined that<br />

the material had an R 2 value of 0,995<br />

for the power regression and that all<br />

of the six data points lay within the<br />

±5% error envelope (Fig. 3). After<br />

assessing the curve prediction analysis,<br />

it was found that at least five of<br />

the data points fell within the ±5%<br />

error envelope using between two<br />

and six samples to obtain the fatigue<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 381


-Coefficient<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

y = −20,428x + 19,535<br />

R 2 = 0,2612<br />

0<br />

0,20 0,25 0,30 0,35 0,40 0,45 0,50 0,55<br />

Fig. 4: L inear regression on b-coefficient versus fiber volume fraction (0/±45 fibers)<br />

Average % of samples within error<br />

100%<br />

95%<br />

90%<br />

85%<br />

80%<br />

75%<br />

70%<br />

65%<br />

60%<br />

55%<br />

50%<br />

Number of samples a-Coefficient b-Coefficient R 2 value Within 5% Within 10%<br />

used<br />

2 5,969 14,245 1 6 6<br />

3 5,998 14,246 1 6 6<br />

4 4,990 14,122 0,998 5 6<br />

5 5,006 14,085 0,997 5 6<br />

6 6,616 14,39 0,995 6 6<br />

Table 2: Model coefficients and prediction analysis for glass/vinyl ester composite<br />

coefficients (Table 2). An appropriate<br />

b- approximation has not yet been<br />

determined for this resin type, so a<br />

single-sample fatigue life prediction<br />

could not be obtained.<br />

Component Data<br />

Generally, few fatigue tests are run<br />

to failure on composite components<br />

owing to the difficulty and expense<br />

of full-scale testing. One researcher 7<br />

ran six bending fatigue tests each on<br />

two glass/vinyl ester composite beams<br />

(one an I-beam, the other a box<br />

beam). Using his data and an adaptation<br />

of U 0 for strain and bending [see<br />

Eq. (12), where I is the moment of<br />

inertia of the sample, L is again the<br />

gage length, and c is the distance from<br />

the neutral axis], a fit and prediction<br />

Fiber volume fraction<br />

Fig. 5: Erro r from varying number of samples used to plot the prediction curve<br />

0,60<br />

5% log error<br />

10% log error<br />

1 2 3 4 5 6 7 8 9<br />

Number of samples used (out of 9)<br />

analysis could be run for full-scale<br />

components. Segments of the beams,<br />

each of dimensions 120 mm × 120 mm<br />

× 6,5 mm were fatigue tested over a<br />

span of 1,83 m.<br />

mean<br />

ILEstatic<br />

U0<br />

= 2<br />

2<br />

(12)<br />

6c<br />

The R 2 values on the power regression<br />

for the box and I-beams were<br />

0,967 and 0,985, respectively. All<br />

12 of the data points lay within the<br />

±5% error envelopes; in fact, they<br />

were also all within curves of ±2,5%<br />

log(N f ) (Fig. 6). The prediction analysis<br />

revealed that using anywhere<br />

between three and six data points<br />

to obtain the fatigue coefficients<br />

resulted in all six data points being<br />

within the ±2,5% error envelope for<br />

both beams. If only two data points<br />

were used to obtain the fatigue coefficients,<br />

five of the six data points<br />

were within the ±2,5% error envelope<br />

and all six were within the<br />

±5% error envelope. An appropriate<br />

b- approximation has not yet been<br />

determined for component fatigue,<br />

so a single-sample fatigue life prediction<br />

could not be obtained.<br />

Environmental Effects on<br />

Fatigue in Composites<br />

Both increased temperature and the<br />

presence of moisture have similar<br />

effects on polymeric composites—both<br />

induce stress by swelling and both relax<br />

stress by softening. 8 Several researchers<br />

9 have shown that the fatigue damage<br />

accumulation sequence was essentially<br />

the same at each temperature level<br />

they tested (ranging between −20 and<br />

150°C) but the rate of accumulation<br />

increased with temperature.<br />

Fatigue tests on the tested glass/vinyl<br />

ester composite coupons were conducted<br />

under varying environmental<br />

conditions in both bending and tension–tension<br />

fatigue with an R-ratio<br />

(s min /s max ) of 0,2. The environmental<br />

conditions included the absence or<br />

presence of salt water during testing,<br />

elevated temperatures, and accelerated<br />

immersion preconditioning (at<br />

12 MPa), and testing was performed<br />

at several stress levels (from 35 to<br />

70% of ultimate stress). While it has<br />

been shown 10 that the absence or<br />

presence of salt often has little effect<br />

on the performance of composites,<br />

salt water was used for the testing in<br />

order to most closely simulate ocean<br />

conditions.<br />

In bending fatigue, the most common<br />

failure mode was delamination of the<br />

outer layers on the tension side of the<br />

sample, believed to be initiated by<br />

microcracking at the midpoint. The<br />

presence of salt water reduced the<br />

fatigue life of the material to 84% of<br />

the value of a dry test at room temperature<br />

and a stress level of 63% of<br />

ultimate load, and to 54% of the value<br />

of a dry test at room temperature and<br />

a stress level of 50%. The effect of<br />

temperature on the life of the material<br />

was found to be approximately linear<br />

(with a higher temperature leading to<br />

a reduction in fatigue life), at least on<br />

those samples that were also tested in<br />

a salt water environment (see Fig. 7).<br />

The tests confirmed that the fatigue<br />

life of the material is decreased by the<br />

382 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Normalized max strain (e max /e ult )<br />

Fig. 6: I-beam fa tigue model fit<br />

Temperature (°C)<br />

0,85<br />

0,80<br />

0,75<br />

0,70<br />

0,65<br />

0,60<br />

0,55<br />

Actual<br />

Prediction<br />

Upper limit (+10%)<br />

Lower limit (−10%)<br />

0,50<br />

Upper limit (+2,5%)<br />

Lower limit (−2,5%)<br />

0,45<br />

4 4,5 5 5,5 6<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Cycles to failure<br />

Linear (cycles to failure)<br />

Fig. 7: Temperature ef fect on bending fatigue<br />

presence of salt water, elevated temperature,<br />

or increased loading level.<br />

The failure mode for all of the tension–tension<br />

fatigue tests seemed to<br />

be a combination of lamina separation<br />

and fiber breaking. The presence of salt<br />

water in the 50% stress level sample<br />

reduced the fatigue life of the material<br />

to 43% of the value of a dry test,<br />

while at 63% no salt water effect was<br />

observed—this makes it clear that salt<br />

water has a larger effect over time, as<br />

the 63% stress level tension test only<br />

lasted for a few hours. As compared to<br />

bending fatigue data under the same<br />

stress levels, the tension fatigue tests<br />

exhibited much fewer cycles to failure<br />

and were more susceptible to environmental<br />

effects. This increased susceptibility<br />

in tension fatigue is likely due<br />

to the higher overall stress on fiber/<br />

resin interfaces and the increased tensioned-surface<br />

area, allowing greater<br />

permeability.<br />

Through the absorption tests, the salt<br />

water saturation point of the material<br />

was determined to be between 0,10<br />

and 0,17% of the weight. The slight<br />

increase in immersion temperature<br />

Cycles to failure (Log N)<br />

0<br />

0 5000 10 000 15 000 20 000 25 000 30 000 35 000<br />

Number of cycles to failure<br />

(16°C) resulted in a decrease in the<br />

number of cycles to failure in all<br />

three tests by 67%. It appears that<br />

room temperature immersion conditioning<br />

could reduce the fatigue<br />

life of the material to 50 to 65% of<br />

the fatigue life of the pre-immersion<br />

material, while 38°C immersion conditioning<br />

could reduce the fatigue<br />

life of the material to 15 to 25%.<br />

More tests need to be run at different<br />

stress levels before the strain<br />

energy model can be applied to the<br />

tests run under different environmental<br />

conditions.<br />

Conclusion<br />

y = −0,0019x + 74,587<br />

R 2 = 0,985<br />

The strain energy fatigue model<br />

appears to provide both a good fit and<br />

a good prediction for the fatigue life<br />

of GFRP composite materials. The<br />

large amount of data analyzed from<br />

the composite material fatigue database<br />

indicates the ability of the model<br />

to fit a variety of GFRP materials with<br />

80 to 90% of the tests falling within<br />

±5% of the log number of cycles to<br />

failure. The model was also shown to<br />

be able to predict the fatigue life of<br />

polyester GFRPs to within ±5% of<br />

the log number of cycles to failure<br />

using only two experimental values<br />

with a success rate of over 75%; using<br />

three increased the success rate to<br />

over 82%. It appears that the model<br />

can predict values nearly as accurately<br />

with nine samples as it can with<br />

only two or three samples, illustrating<br />

how it only requires a minimum<br />

number of experimental data points.<br />

The results of the analysis of the component<br />

fatigue tests suggest that the<br />

model is able to fit and predict the<br />

fatigue life for a full-scale composite<br />

component as accurately as for a coupon<br />

while maintaining the same level<br />

of simplicity. The model was able to<br />

fit and predict the fatigue life of both<br />

of the components tested to within<br />

±2,5% of the log number of cycles to<br />

failure.<br />

The introduction of moisture or elevated<br />

temperatures has been shown<br />

to significantly decrease the fatigue<br />

life of the glass/vinyl ester composite<br />

material. This effect seems to be particularly<br />

linear for temperature changes<br />

and is relative to the testing length for<br />

the presence of salt water. Pre-immersion<br />

appears to produce a consistent<br />

reduction in fatigue life, depending on<br />

the immersion temperature.<br />

The strain energy fatigue model provides<br />

a simple method to predict<br />

fatigue life to within acceptable levels<br />

of accuracy in many structures, such as<br />

windmill blades, bridge decks, or deepsea<br />

pipes. Additional work should<br />

focus on developing material and testing<br />

condition-based approximations<br />

for b-coefficients, as well as continuing<br />

to test the model against componentscale<br />

and environmental fatigue test<br />

results.<br />

References<br />

[1] Liang R, GangaRao HV. Applications of<br />

fiber reinforced polymer composites. In Polymer<br />

Composites III, 2004 Creese R, GangaRao<br />

HV (eds). DEStech: Lancaster (PA), 2004;<br />

173–187.<br />

[2] Samborsky DD, Mandell JF. DOE/MSU<br />

Composite Material Fatigue Database, Version<br />

18.1, 2005.<br />

[3] Mandell JF, Samborsky DD. DOE/MSU<br />

Composite Material Fatigue Database: Test<br />

Methods, Materials, and Analysis. Contractor<br />

Report SAND97-3002, Albuquerque, NM:<br />

Sandia National Laboratories, 1997.<br />

[4] Natarajan V, GangaRao HV, Shekar V.<br />

Fatigue response of fabric-reinforced polymeric<br />

composites. J. Compos. Mater. 2005; 39(17):<br />

1541–1559.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 383


[5] Ellyin F, El-Kadi H. A fatigue failure criterion<br />

for fiber reinforced composite laminae.<br />

Composite Structure 1990; 15: 61–74.<br />

[6] Dittenber DB, GangaRao HV. Evaluation of<br />

fatigue life prediction model for GFRP composite<br />

materials. Proceedings of SPE-ANTEC <strong>2010</strong>;<br />

<strong>2010</strong> May 16–20; Orlando (FL), United States;<br />

<strong>2010</strong>.<br />

[7] Nagaraj V. Static and Fatigue Response of<br />

Pultruded FRP Beams without and with Splice<br />

Connections. College of Engineering and<br />

Mineral Resources [thesis]. Morgantown (WV),<br />

West Virginia University, 1994.<br />

[8] Weitsman Y. Moisture in composites: sorption<br />

and damage. In Fatigue of Composite Materials.<br />

Reifsnider, KL (ed.). Elsevier: Amsterdam, 1990;<br />

385–429.<br />

[9] Khan R, Khan Z, Al-Sulaiman F, Merah<br />

N. Fatigue life estimates in woven carbon<br />

fabric/epoxy composites at non-ambient temperatures.<br />

J. Compos. Mater. 2002; 36(22):<br />

2517–2535.<br />

[10] McBagonluri F, Garcia K, Hayes M,<br />

Verghese KNE, Lesko JJ. Characterization of<br />

fatigue and combined environment on durability<br />

performance of glass/vinyl ester composite<br />

for infrastructure applications. Int. J. Fatigue<br />

1999; 22: 53–64.<br />

Further Information<br />

1. Constructed Facilities Center at<br />

West Virginia University. http://www.<br />

cemr.wvu.edu/cfc/.<br />

2. Department of Energy/Montana<br />

State University Composite Material<br />

Fatigue Database. http://www.coe.<br />

montana.edu/composites/.<br />

384 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


A Composite Bridge is Favoured by Quantifying<br />

Ecological Impact<br />

Ryszard A. Daniel, PhD, Eng., Ministry of Transport, Public Works and Water Management, Division of Infrastructure, Utrecht,<br />

The Netherlands. Contact: richard.daniel@rws.nl<br />

Abstract<br />

Carrying traffic loads is not the only objective of bridge designers nowadays.<br />

Other demands include constructing a bridge in a sustainable way, which reduces<br />

pollution and other harm to the environment. In The Netherlands, the government<br />

responds to such demands by promoting technologies and materials that<br />

decrease the environmental impact of construction projects.<br />

An assessment of that impact is, however, quite complex for bridge projects.<br />

The existing analytical methods, such as life-cycle analysis (LCA), require an<br />

extensive data input. Moreover, their results are more reliable for relatively<br />

simple products of short life cycles, for example, door or window frames, than<br />

for complex construction projects. In construction projects, the life cycles cannot<br />

be determined with the same precision and the materials are usually chosen<br />

in the very early stage of design. As a result, the data required by the LCA are<br />

often incomplete or even disputable. Therefore, there is a demand for ecological<br />

analysis methods that enable quick scanning of several material options, require<br />

less-extensive data input and are hardly, or not, vulnerable to arbitrariness.<br />

Keywords: FRP structures; eco-analysis, material choice; sustainable material;<br />

sustainable bridge; energy input; exergy; emissions; pollution data.<br />

This paper answers the above-mentioned<br />

demand by presenting a method<br />

for ecological material selection for<br />

a bridge. It shows a way to quantify<br />

the environmental impacts of possible<br />

material choices in comparable<br />

terms and to assess those choices with<br />

respect to their impact. The method<br />

was first developed and applied for<br />

the quay footbridges in the Noordland<br />

inner harbour, province of Zeeland,<br />

The Netherlands. Five material options<br />

were considered: structural steel, stainless<br />

steel, composites (fibre-reinforced<br />

polymers, FRPs), aluminium and reinforced<br />

concrete. The analysis allowed<br />

evaluating these options in terms of<br />

three crucial ecological indicators:<br />

energy consumption, pollution to air<br />

and pollution to water.<br />

The ecological analysis was performed<br />

along with the costs and<br />

service-life assessment. The comperforming<br />

remarkably well since<br />

then, validating the computed ecological<br />

and other indicators. Its good<br />

performance suggests the possible<br />

construction of more similar footbridges<br />

in that area in future. This<br />

paper presents a comparison of those<br />

indicators for the material options<br />

considered, and discusses these and<br />

some selected problems of the ecological<br />

analyses.<br />

The applied ecological analysis has<br />

been presented on various occasions<br />

since the bridge construction. 1–3 Yet,<br />

it still evokes much interest because<br />

of the importance of environmental<br />

engineering in relation to, for example,<br />

climatic processes. This paper aims to<br />

respond to that interest, giving more<br />

details of both the applied ecological<br />

analysis and the constructed FRP<br />

bridge.<br />

Introduction<br />

puted performances of all the material<br />

options considered resulted in<br />

an advice to construct a bridge of<br />

pultruded FRP profiles (Fig. 1). The<br />

customer followed that advice. It was<br />

the first bridge constructed using this<br />

technology in The Netherlands. The<br />

bridge was assembled and brought<br />

into service in 2001. It has been<br />

Project Objectives and Scope<br />

The Dutch province of Zeeland is<br />

a costal area in the south-western<br />

delta of the rivers Rhine, Meuse and<br />

Scheldt. High exposure to sea water,<br />

wind loads and chloride corrosion<br />

form part of the usual design specifications.<br />

At the end of 1999, the Regional<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: January 14, <strong>2010</strong><br />

Paper accepted: July 19, <strong>2010</strong><br />

Fig. 1: Installation of the Noordland inner harbour footbridge<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 385


North<br />

Sea<br />

B<br />

NL<br />

Authority for Public Works and Water<br />

Management ordered an investigation<br />

on construction materials for a footbridge<br />

in the Noordland inner harbour<br />

that forms part of the Eastern Scheldt<br />

Storm Surge Barrier complex. The<br />

bridge provides a double span access<br />

to a mooring pontoon (Fig. 2). The<br />

new bridge was to replace the old steel<br />

bridge that was largely deteriorated by<br />

corrosion after only 35 years of service.<br />

This was not surprising, considering the<br />

extreme conditions at that location.<br />

The service load of the bridge is 400<br />

kN/m 2 . Other loads are wind, snow,<br />

glitter ice, and so on. There is no navigation<br />

under the bridge. The support<br />

level to pontoon varies because of the<br />

tides. The allowable span deflection<br />

is limited to 1/250. The customer was<br />

interested in comparing the performances<br />

of the first four bridge materials<br />

from the following list:<br />

– structural steel (with coating);<br />

– stainless steel;<br />

– synthetic material (composite);<br />

– aluminium;<br />

– concrete.<br />

The fifth material was investigated<br />

later for the sake of completeness. The<br />

weight of a concrete bridge made it<br />

unfit for a pontoon support. Timber<br />

was also not an interesting option<br />

because of its maintenance requirements,<br />

combustibility and short service<br />

life at this particular location.<br />

Nonetheless, it certainly can be considered—also<br />

with respect to the environment—in<br />

other bridge projects. In this<br />

paper, timber is not included, because<br />

the considerations that determine its<br />

environmental performances are of a<br />

different nature. An important criterion<br />

is, for example, sustainable forest<br />

management. 4 It is difficult to quantify<br />

such criteria in a manner that allows<br />

for a comparison with other materials.<br />

The performances of each option had to<br />

be quantified in terms of the following<br />

D<br />

A<br />

A<br />

13,5<br />

Fig. 2: Bridge location and dimensions (Units: m)<br />

13,5<br />

Pontoon<br />

1,60<br />

A - A<br />

1,10<br />

four criteria: construction costs, maintenance<br />

costs, service life and environmental<br />

impact. Aesthetics was not a<br />

prior concern at this desolate location.<br />

Maintenance and service life appeared<br />

to show a strong correlation. It was,<br />

therefore, agreed to impose a uniform<br />

service life of 50 years on all material<br />

options. This period reflects the current<br />

design views in The Netherlands.<br />

In this way, the number of assessment<br />

criteria was reduced to three, which<br />

simplified the analysis.<br />

Construction and maintenance costs<br />

are quite common criteria in engineering;<br />

therefore, only the final results are<br />

presented. To quantify the environmental<br />

impact, however, an investigation<br />

method had to be set up first. As<br />

already discussed, the existing methods<br />

like the LCA 5 were not very helpful.<br />

The footbridge appeared to be too<br />

complex and too vaguely determined<br />

at this stage. Making detailed bridge<br />

designs and life-cycle inventories for<br />

all material options was, obviously, not<br />

the intention. Therefore, a simplified,<br />

but workable, two-way evaluation was<br />

chosen:<br />

– energy consumption analysis—<br />

taking also account of the energy<br />

“stored” in materials and products<br />

(the so-called “exergy” method 6 );<br />

– analysis of loads (pollutions) to water<br />

and air as a result of material winning,<br />

processing, fabrication of the<br />

final product and its maintenance.<br />

In current views, the first approach<br />

can be seen as a measure of not only<br />

energy consumption as such (i.e.<br />

decrease of global energy resources)<br />

but also the processes resulting from<br />

fossil fuel combustion, like the greenhouse<br />

effect, rise in ocean level, global<br />

climatic changes, and so on. The second<br />

approach (loads to air and water<br />

apart) produced global pollution<br />

data of the bridge options under consideration.<br />

Loads to soil appeared<br />

to be insignificant, but they can be<br />

analysed in the same way, whenever<br />

relevant.<br />

Conceptual Designs<br />

As the materials in question represented<br />

in fact five groups of materials,<br />

the material grades had to be chosen. In<br />

accordance with the existing practice,<br />

the following grades were selected:<br />

– structural steel: S235J0 or S355J0,<br />

according to the European norm<br />

EN 10025. An arc-sprayed aluminium<br />

coating was considered as an<br />

alternative to the conventional paint<br />

system.<br />

– stainless steel: X2CrNi18-11 or<br />

X2CrNiMo18-14-3 according to the<br />

European norms (AISI 304L or 316L<br />

according to the US standards).<br />

– composite: fibreglass-reinforced<br />

polyester resin (FGRP) in pultruded<br />

sections.<br />

– aluminium: AlMgSi1,0 F31 according<br />

to the DIN 1748 code (or 6061<br />

and 6063 alloys according to the<br />

ASTM B221).<br />

– concrete: B35 according to the<br />

European norm EN 1992-1, 150 kg<br />

of reinforcement per 1 m 3 ; 100 kg<br />

of other steel accessories (e.g. handrails)<br />

per 1 m 3 .<br />

The next step was to complete five<br />

rough conceptual bridge designs, one<br />

in each optional material. It soon<br />

became clear that each option required<br />

a different form, system, manufacturing<br />

approach, and so on. In structural<br />

steel and concrete, for example, conventional<br />

girders with separate handrails<br />

were an evident choice, whereas<br />

in the other, more expensive materials<br />

the handrails were integrated in truss<br />

or truss-like girders. Major differences<br />

appeared also in section shapes, deck<br />

systems, and so on. In Fig. 3, one span<br />

of the bridge in each of the five materials<br />

is shown. The structural analysis was<br />

very brief in all cases. Nevertheless, it is<br />

fair to say that the bridge spans shown<br />

in Fig. 3 are representative for the<br />

considered materials, and comparable<br />

with each other in terms of strength<br />

and durability.<br />

The material mass estimations are<br />

based on a brief analysis and data<br />

from similar projects. These masses<br />

form the data for estimating both<br />

total costs and environmental impact.<br />

Remarkably large mass differences are<br />

seen between the material options. This<br />

requires a few comments. The mass of<br />

structural steel span would have been<br />

lower (2200–2500 kg) if truss girders<br />

386 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Girder h = 360<br />

of HE240A<br />

Lower chord<br />

φ168, 3 × 6,3<br />

Upper chord/Handrail<br />

Lower chord<br />

U240 × 72 × 8<br />

Diagonals<br />

φ 70 × 60 × 5<br />

Chords<br />

φ 200 × 100 × 10<br />

Girder<br />

φ 450 × 200<br />

Option 1: Structural steel<br />

One span mass: 3000 kg<br />

Cross girders<br />

HE160B<br />

Option 2: Stainless steel<br />

One span mass: 2800 kg<br />

Cross girders<br />

φ139, 7 × 6,3<br />

Option 3: Composite<br />

One span mass: 2000 kg<br />

Cross girders<br />

H200 × 100 × 10<br />

Option 4: Aluminium<br />

One span mass: 1600 kg<br />

Option 5: Concrete<br />

One span mass: 14 000 kg<br />

Deck plate 150<br />

FSC timber 50 mm<br />

Cross girder<br />

HE240B<br />

FSC timber 50 mm<br />

Cross girder<br />

φ168, 3 × 6,3<br />

Cross girder<br />

13 500 1600<br />

Fig. 3: Bridge span in five material options (length units: mm)<br />

integrated with handrails were chosen<br />

instead of beams. This has deliberately<br />

not been done to justify neglecting<br />

the impact of steel coating. In any<br />

case, however, the composite and aluminium<br />

bridges appear to be the lightest.<br />

The concrete bridge is 5–10 times<br />

heavier than the other bridges. The<br />

dead weight was of minor importance<br />

here, as long as it did not cause pontoon<br />

overloading. A smaller weight is,<br />

however, desirable in large bridges. It<br />

allows for higher traffic loads, lighter<br />

foundations, pillars, transport and<br />

assembly equipment.<br />

Global Assessment<br />

1100<br />

The bridge conceptual designs were<br />

employed to collect more data for the<br />

evaluation—not only the total material<br />

masses. The drawings in the form<br />

of outlines prompted specific questions<br />

and enabled collection of relevant<br />

data on the market. The desired<br />

data covered, in general, the following<br />

subjects:<br />

– quantities and unit prices of the<br />

materials involved;<br />

– available manufacturing technologies,<br />

their costs, conditions, quality<br />

assurances and risks;<br />

– transport and assembly requirements,<br />

like access, time, heavy equipment,<br />

specific provisions;<br />

– inspection and maintenance frequencies<br />

during the service life;<br />

– environmental impact of all processes<br />

involved.<br />

The accuracy of these data was not<br />

always high because of the preliminary<br />

nature of bridge design. In some cases,<br />

rough estimations had to be made. The<br />

concerned specialists agreed, nonetheless,<br />

that a sufficient, well balanced base<br />

was provided to evaluate the bridge<br />

options. The concise results of this<br />

evaluation are shown in Table 1. The<br />

general conclusions are as follows:<br />

– In terms of construction costs, the<br />

structural steel and the concrete<br />

bridges are favourable. The stainless<br />

steel bridge is too expensive; the<br />

composite and aluminium bridges<br />

score in the middle.<br />

– In terms of maintenance costs, the<br />

scores are opposite. The stainless<br />

steel bridge is the cheapest, followed<br />

by the concrete bridge. The structural<br />

steel bridge (conventionally<br />

painted) is the most expensive. The<br />

scores of the composite and aluminium<br />

bridges lie in between.<br />

– Adding construction and maintenance<br />

costs (whether or not capitalized)<br />

puts the concrete bridge in the<br />

first place and the structural steel<br />

bridge in the second. The composite<br />

bridge takes a good third place,<br />

closely followed by aluminium. The<br />

stainless steel bridge is evidently the<br />

most expensive.<br />

– Analysis of the energy consumption<br />

makes the composite bridge a<br />

winner. Every other option results<br />

in energy consumption that is more<br />

than two times as high. Energy consumption<br />

is seen as an important<br />

indicator of the contribution to the<br />

global warming effect.<br />

– The composite bridge is also the best<br />

in terms of the resulting water and<br />

air pollution levels. The structural<br />

steel bridge is the second, concrete<br />

bridge is the third and aluminium<br />

bridge is the fourth.<br />

The customer was advised as follows:<br />

if construction cost was the primary<br />

concern, the choice of a structural<br />

steel bridge was the best. But if a little<br />

extra cost was acceptable in the interest<br />

of the environment, the composite<br />

bridge of pultruded profiles was the<br />

best choice. An additional argument<br />

in support of the composite bridge<br />

was the innovative character of such<br />

a project. It was to be the first composite<br />

bridge of pultruded profiles in<br />

The Netherlands. The customer was<br />

indeed in a position, and willing, to<br />

choose the second, pro-environmental<br />

option. The composite bridge was constructed<br />

in October 2001. It has been<br />

closely monitored since then, confirming<br />

the results of the analysis.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 387


Bridge material<br />

Construction costs<br />

(EUR)<br />

Maintenance costs<br />

(EUR)<br />

Criterion<br />

Environment: Energy<br />

consumption (MJ)<br />

Environment: Critical volume<br />

of polluted air and water<br />

Structural steel Painted: 40 000 Painted: 30 000 “Exergy” method: 294 000 Water: 697,4 m 3<br />

Aluminium coated: 50 000 Aluminium coated: 6 000 Air: 7,09 × 10 6 m 3<br />

Stainless steel Steel AISI 316L: 110 000<br />

Steel AISI 304L: 96 000<br />

Steel AISI 316L: 6000<br />

AISI 304L more, life<br />

cycle shorter<br />

Rough estimation:<br />

“Exergy” method: 329 600 Not investigated but certainly<br />

more pollution than for structural<br />

steel<br />

“Exergy” method: 120 000 Water: 85,8 m 3<br />

Composite Pultruded sections of<br />

FGRP: 70 000<br />

17 000<br />

Air: 7,92 × 10 6 m 3<br />

Aluminium Quality AlMgSi1 Rough estimation: 19 000 “Exergy” method: 268 700 Water: 565,3 m 3<br />

acc. to DIN 1748: 77 000 Air: 41,10 × 10 6 m 3<br />

Concrete Reinforced concrete B35,<br />

handrails etc: 30 000<br />

Rough estimation:<br />

10 000<br />

Table 1: Performances of the five material options for the bridge<br />

“Exergy” method: 277 200 Water: 341,9 m 3<br />

Air: 31,04 × 10 6 m 3<br />

Eco-Analysis in Terms<br />

of Energy Consumption<br />

Ecological performances of a particular<br />

material option cannot be expressed in<br />

a single indicator, although it is advisable<br />

to keep the number of indicators<br />

small. Energy consumption, therefore,<br />

does not reveal everything about the<br />

ecological performances, but it is an<br />

important indicator in this field. It<br />

requires no argument today that energy<br />

consumption is a global environmental<br />

issue in both direct and indirect senses.<br />

In the first sense, it decreases the global<br />

energy resources which are—for the<br />

biggest part—not renewable. In the<br />

second sense, it harms the environment<br />

in many ways, including its contribution<br />

to the emission of CO 2 , other<br />

“greenhouse gases” and the resulting<br />

climatic changes. However, if the latter<br />

is seen as the main or only issue of ecoanalysis<br />

(which is not the author’s view),<br />

a direct analysis of greenhouse gas<br />

emission, 7 will be more appropriate.<br />

The required data is that of the energy<br />

use for the processing and manufacturing—from<br />

obtaining the raw materials<br />

to the final product—of one mass<br />

unit of the product in question (in<br />

MJ/kg). These data vary because the<br />

same materials and products can be<br />

obtained using different technologies.<br />

As eco-analyses are quite new, there is<br />

still much arbitrariness in defining the<br />

data. Therefore, it is always advisable<br />

to check which processes are covered<br />

by the received data. During this study,<br />

for example, the following energy consumption<br />

rates for structural steel products<br />

were found in various sources:<br />

– Source 1 (The Netherlands): 6<br />

46 MJ/kg;<br />

– Source 2 (The Netherlands): 8<br />

31 MJ/kg;<br />

– Source 3 (The Netherlands): 9<br />

18 MJ/kg;<br />

– Source 4 (USA): 10 6 MJ/kg.<br />

Such differences may be surprising to<br />

engineers who are used to approved<br />

specifications, standard codes and reliable<br />

and well tested data. However,<br />

the databases held by various institutes<br />

appear to be usable. When high figures,<br />

for example, for structural steel are<br />

quoted, they usually include energy<br />

input for rolling, surface treatment,<br />

transport, welding, fabrication, delivery<br />

and assembly of the structure. Low<br />

figures comprise smaller numbers of<br />

those processes. Data on other materials<br />

are collected in a similar way so that<br />

every database is usually consistent. It<br />

is, therefore, recommended to use data<br />

from the same source throughout the<br />

entire analysis. The lack of standards<br />

should temporarily be accepted. In<br />

the interest of the environment, one<br />

should rather critically apply the existing<br />

data than wait until they become<br />

better.<br />

The so-called “exergy” method was<br />

used to quantify the energy use for the<br />

five bridge options. In this method, the<br />

total energy consumption is a sum of<br />

energy value decreases for the materials<br />

in the processes involved. The<br />

analysis was limited to basic materials;<br />

wooden bridge decks in both structural<br />

and stainless steel bridges, stainless<br />

steel connectors in aluminium and<br />

FRP bridges and so on were ignored.<br />

The energy consumption rates per<br />

material unit were collected from the<br />

first 6 database except for composites<br />

(second 8 data base). Although both<br />

companies were involved in the official<br />

“Eco-indicator” project, 11 no uniform<br />

energy database for all materials<br />

was available at that time. The review<br />

resulted in some adjustments to the<br />

data for the purpose of this analysis<br />

(Table 2). According to recent views,<br />

the data for composites might still<br />

require a minor increase. These data<br />

should, however, not be confused with<br />

the much higher energy rates for plastics.<br />

Polyester resin usually makes up<br />

less than 50% in volume (about 30%<br />

in weight) of pultruded profiles. The<br />

rest is fibreglass.<br />

In the following example, energy consumption<br />

is estimated for a structural<br />

steel bridge:<br />

Total mass of two spans: 6000 kg.<br />

Assumed: 80% of the primary and<br />

20% of the secondary (recycled)<br />

material. Energy consumption on<br />

delivery:<br />

Ex 0 = 6000 × [0,8 × (46–7) + 0,2<br />

(1)<br />

× (36–7)] = 222 000 MJ<br />

The energy used during maintenance<br />

(2 × blast cleaning and painting) was<br />

approximated by subtracting the figure<br />

for unpainted structure (31 MJ/kg)<br />

obtained from another database. 9 To<br />

take account of the time delay (about<br />

20 and 35 years), a factor of 0,75 was<br />

introduced:<br />

Ex 1 = 6000 × 2 × 0,75 × (46 − 7 − 31)<br />

= 72 000 MJ (2)<br />

This gives the total energy<br />

consumption:<br />

Ex = Ex + Ex = 222 000 + 72 000<br />

0 1<br />

= 294 000 MJ<br />

(3)<br />

The energy consumptions for the other<br />

material options were estimated in a<br />

similar manner. This gave the energy<br />

impact graph for all the five bridge<br />

options (Fig. 4).<br />

388 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Material<br />

Structural steel<br />

(e.g. S235J0)<br />

Stainless steel<br />

(e.g. AISI 316L)<br />

Composites<br />

(FGRP)<br />

Aluminium<br />

(e.g. AlMgSi1)<br />

Reinforced concrete<br />

(B35, handrails)<br />

Condition<br />

Primary<br />

Secondary<br />

Primary<br />

Secondary<br />

Primary<br />

Secondary<br />

These results are not as “hard” as, for<br />

example, those from structural analyses.<br />

One may wonder why the delay factor<br />

of 0,75 is used for the maintenance<br />

of the structural steel bridge—and if<br />

so, then why it is not applied to deck<br />

replacements in other bridge options.<br />

In this case, the engineers felt that spare<br />

decks of “unusual” materials (composite,<br />

aluminium) should be secured, that<br />

is, delivered together with the bridges.<br />

This assumption is, however, arbitrary.<br />

Another simplification is that the<br />

energy for dismantling after the service<br />

life has been neglected. Including<br />

it would probably point to the concrete<br />

bridge as the most energy-consuming<br />

option. Concrete demolition<br />

and utilization requires much energy.<br />

As mentioned, there are also differences<br />

in energy rating between various<br />

institutions and countries, especially in<br />

regard to composites. German data, 12<br />

often result in higher energy rates<br />

and American data 10 in lower energy<br />

rates. However, it is undisputable that<br />

the composite bridge had the lowest<br />

energy consumption.<br />

Loads to the Environment<br />

Energy analyses do not indicate how<br />

“clean” or “dirty” the considered<br />

Energy consumption<br />

value (MJ/kg)<br />

46<br />

36<br />

69<br />

54<br />

33<br />

—<br />

Remaining “stored”<br />

energy (MJ/kg)<br />

options are, that is, they provide no<br />

comparison in terms of environmental<br />

pollution. The problem with such a<br />

comparison is that each material option<br />

gives a spectrum of qualitatively different<br />

pollutions, which cannot simply<br />

be added up. The solution is found by<br />

taking account of the so-called “legal<br />

thresholds” of the particular pollutants.<br />

This was, to the author’s best knowledge,<br />

the first time that this approach<br />

was used in an infrastructure project.<br />

The applied method is derived from<br />

the so-called critical load method, 10<br />

and is based on the following two data<br />

records:<br />

– B m,i (kg/m 3 ), emitted masses of the<br />

pollutants i due to production and<br />

processing of 1 m 3 of the material m.<br />

Such emissions are usually recorded<br />

as loads to air, water and (exceptionally)<br />

soil.<br />

– B 0,i (kg/m 3 ), legal thresholds of the<br />

pollutants i in 1 m 3 of air, water and<br />

(exceptionally) soil.<br />

When these two data records are<br />

known along with the total mass G m<br />

and density γ m of the material m, the<br />

total critical volume of polluted air<br />

V a m or water V w m (m 3 ) can be computed<br />

as follows:<br />

7<br />

7<br />

11<br />

11<br />

9<br />

—<br />

Primary 137 33<br />

Secondary 45 33<br />

Primary<br />

Secondary<br />

Table 2: Energy consumption data for the five material options for the bridge<br />

350 000<br />

300 000<br />

250 000<br />

200 000<br />

150 000<br />

100 000<br />

50 000<br />

0<br />

Structural<br />

steel<br />

On delivery<br />

Stainless<br />

steel<br />

11<br />

—<br />

Maintenance<br />

Composite Aluminium Reinforced<br />

concrete<br />

Fig. 4: Energy impact of the bridge for the five material options<br />

2<br />

—<br />

V<br />

G<br />

γ<br />

m<br />

m<br />

= ×∑<br />

m i<br />

B<br />

B<br />

mi ,<br />

0 , i<br />

(4)<br />

Tables 3 and 4 present the emissions<br />

B m,i and their legal thresholds B 0,i for<br />

the four final material options: structural<br />

steel, composite, aluminium and<br />

concrete. The stainless steel option was<br />

not given up at that stage. The data for<br />

structural steel and aluminium bridges<br />

were collected from Refs. [10, 13]. The<br />

emission data for polyester resin were<br />

provided by the world market leader<br />

in this branch, and combined with the<br />

data for glass to give the aggregated<br />

emissions for FRP. The data for reinforced<br />

concrete (including steel accessories<br />

like handrails) were obtained<br />

by combining the records for concrete<br />

and steel.<br />

Apart from the global results (see<br />

Table 1), it is interesting to compare<br />

the pollutions to water and air qualitatively.<br />

For example, for the composite<br />

bridge, Eq. (4) and the data in<br />

Tables 3 and 4 give the following critical<br />

volumes of polluted air, V a cp and<br />

water V w cp:<br />

V<br />

V<br />

a<br />

cp<br />

w<br />

cp<br />

G B<br />

cp<br />

cp,<br />

i<br />

= × ∑ =<br />

γ i B<br />

cp<br />

0, i<br />

3<br />

−1<br />

4000 ⎛ 103 , × 10 12 , × 10 ⎞<br />

= ×<br />

+ ⋅⋅⋅+<br />

3<br />

7<br />

1700 ⎝<br />

⎜<br />

−<br />

−<br />

90 , × 10 80 , × 10 ⎠<br />

⎟<br />

= 235 , × 3, 37 × 10 6 = 7, 92 × 10<br />

6 m 3<br />

(5)<br />

G B<br />

cp<br />

cp,<br />

i<br />

= × ∑ =<br />

γ B<br />

cp<br />

i<br />

0, i<br />

−6<br />

−2<br />

4000 ⎛ 20 , × 10 30 , × 10 ⎞<br />

= ×<br />

−5<br />

−3<br />

1700<br />

⎜ + ⋅⋅⋅+<br />

⎝ 50 , × 10 10 , × 10<br />

⎟<br />

⎠<br />

= 2, 35 × 36, 5 = 85, 8 m 3 (6)<br />

The components of these sums, multiplied<br />

by the ratio G cp /γ cp , are represented<br />

in diagrams (left) in Fig. 5, along<br />

with the results for the other material<br />

options. The total critical volumes of<br />

polluted air and water are compared<br />

in pie charts (right) in Fig. 5. Also, the<br />

composite bridge appears to be more<br />

favourable than the other considered<br />

options.<br />

The analysis in this paper was deliberately<br />

kept simple. The bridge options<br />

were approached as single-material<br />

cases. Although there usually exists<br />

a single dominant material in all<br />

bridge projects, it may be advisable<br />

to consider other component materials<br />

as well. Examples are concrete<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 389


Polluter Structural steel B st,i Composite B cp,i Aluminium B al,i Concrete B cr,i Threshold B 0,i<br />

CO 2 2,56 × 10 +3 1,03 × 10 +3 2,1 × 10 +4 4,95 × 10 +2 9 × 10 −3<br />

CO 9,58 × 10 +1 1,32 5,15 × 10 +1 3,48 4 × 10 −5<br />

CH 4 5,95 1,21 5,39 × 10 +1 9,89 × 10 −1 6,7 × 10 −3<br />

N 2 O 3,7 × 10 −2 4,8 × 10 −3 2,94 × 10 −1 1,51 × 10 −2 1 × 10 −7<br />

PM Fe/Al-oxi. * 2,2 × 10 −1 1,05 × 10 −1 1,65 6 × 10 −2 1 × 10 −7<br />

PM Si/Ca-oxi. * 4,2 × 10 −2 5,05 × 10 −1 2,7 × 10 −1 4,7 × 10 −1 3 × 10 −7<br />

SO 2 3,28 2,51 × 10 −3 1,27 × 10 +1 2,8 × 10 −1 1,2 × 10 −6<br />

NO x 3,08 2,83 2,45 × 10 +1 1,27 1 × 10 −5<br />

Styrene — 1,2 × 10 −1 — — 8 × 10 −7<br />

*PM = particulate matter (dust), here predominately Fe/Al or Si/Ca oxides.<br />

Table 3: Emissions to air for structural steel, composite, aluminium and reinforced concrete<br />

Polluter Structural steel B st,i Composite B cp,i Aluminium B al,i Concrete B cr,i Threshold B 0,i<br />

Aluminium 3,33 × 10 −6 2 × 10 −6 3,09 × 10 −5 1,65 × 10 −7 5 × 10 −5<br />

Ammonia 4,58 × 10 −3 1,1 × 10 −3 4,23 × 10 −2 2,38 × 10 −4 2,2 × 10 −3<br />

Cadmium 4,57 × 10 −5 2,1 × 10 −6 4,28 × 10 −4 2,18 × 10 −6 3,5 × 10 −6<br />

Copper 1,96 × 10 −8 7,9 × 10 −4 1,82 × 10 −7 0,99 × 10 −9 2 × 10 −4<br />

Cyanide 3,08 × 10 −4 7,4 × 10 −5 2,85 × 10 −3 1,6 × 10 −5 1 × 10 −4<br />

Fluoride 1,03 × 10 −1 2 × 10 −4 6,49 × 10 −3 3,51 × 10 −3 1,5 × 10 −3<br />

Manganese 6,07 × 10 −6 3,6 × 10 −6 5,64 × 10 −5 3,03 × 10 −7 5 × 10 −5<br />

Mercury 1,57 × 10 −4 7 × 10 −7 1,45 × 10 −3 7,53 × 10 −6 5 × 10 −6<br />

Zinc 3,97 1,4 × 10 −3 5,44 × 10 −2 1,35 × 10 −1 5 × 10 −3<br />

Cobalt — 3 × 10 −2 — — 1 × 10 −3<br />

Table 4: Emissions to water for structural steel, composite, aluminium and reinforced concrete<br />

Styrene<br />

NO x<br />

SO 2<br />

PM Si/Ca<br />

PM Fe/Al<br />

N 2 O<br />

Struct. steel<br />

CH 4<br />

Composite<br />

CO<br />

CO 2<br />

in 10 3 m 3 of air Aluminium<br />

Reinf. concrete<br />

0 5000 10 000 15 000 20 000 25 000<br />

Cobalt<br />

Zinc<br />

Mercury<br />

Manganese<br />

Fluoride<br />

Cyanide<br />

Copper<br />

Cadmium<br />

Ammonia<br />

Aluminium<br />

in 10 3 of water<br />

Struct. steel<br />

Composite<br />

Aluminium<br />

Reinf. concrete<br />

0 100 200 300 400 500 600 700<br />

31 040<br />

Loads to air<br />

341,9<br />

565,3<br />

7090<br />

Loads to water<br />

85,8<br />

7920<br />

41 100<br />

Fig. 5: Polluted air and water as a result of bridge construction with four material options<br />

and steel in cable-stayed bridges or<br />

steel and composite in steel bridges<br />

with composite decks. The discussed<br />

697, 4<br />

method can be applied in such cases<br />

too. Equation (4) then takes the following<br />

form:<br />

G B<br />

V = ⎛<br />

j<br />

×<br />

⎞<br />

ji ,<br />

complex ∑⎜<br />

⎟<br />

j ⎝ γ<br />

∑<br />

B<br />

(7)<br />

j i 0, i ⎠<br />

where V complex is the critical volume<br />

of air or water polluted up to the<br />

respective legal threshold (m 3 ); G j<br />

is the total mass of material j in the<br />

considered complex material bridge<br />

option (kg); g j is the specific mass<br />

of material j (kg/m 3 ); B j,i is the mass<br />

of pollutant i emitted by production<br />

+ processing of 1 m 3 of material<br />

j (kg/m 3 ); B 0,i is the respective legal<br />

threshold of pollutant i in air or water<br />

(kg/m 3 ).<br />

This may look complex here, but once<br />

we have the databases B j,i and B 0,i in<br />

a PC, this sum presents no problem.<br />

In fact, it can easily be generated in a<br />

simple spread sheet, along with proper<br />

graphs.<br />

Conclusion and Future Outlook<br />

The considered case proves that synthetic<br />

composites (FRPs) constitute<br />

a very interesting material option for<br />

bridges in terms of environmental<br />

impact. A composite bridge project<br />

390 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Fig. 6: Closed ‘traffic ducts’ concept<br />

requires less than half of the energy<br />

input that is required for an equivalent<br />

project constructed using steel,<br />

stainless steel, aluminium or concrete.<br />

In terms of loads to air, the composite<br />

bridge is the second “cleanest” option<br />

after the steel bridge. In terms of loads<br />

to water, the composite bridge is the<br />

undisputable winner. This makes composites<br />

an advantageous material for<br />

bridges, despite the slightly higher construction<br />

costs.<br />

The main reasons for the good performance<br />

of FRP are:<br />

– good mechanical properties, particularly<br />

the tensile strength, resulting in<br />

small quantities required;<br />

– very good chemical stability, resulting<br />

in low maintenance and long service<br />

life;<br />

– well-controlled processes, resulting<br />

in small error margins and low environment<br />

impact.<br />

The presented case should be seen as<br />

an indication, but not necessarily as<br />

evidence, for other bridge projects.<br />

Individual requirements and local<br />

conditions often play a decisive role<br />

in material selection. In the considered<br />

Noordland Bridge, for example,<br />

high corrosion resistance was particularly<br />

valued because of the surrounding<br />

environment (sea water). For road<br />

bridges, the relatively low elasticity<br />

modulus of composites may limit their<br />

applications or require other forms and<br />

structural systems, for example, “ribbon<br />

bridge”, 14 membrane deck, 15 high truss<br />

girders or closed traffic ducts. 16 The latter<br />

also (Fig. 6) offer other advantages<br />

for the environment. Yet, as the significance<br />

of environmental performances<br />

steadily grows, the synthetic composites<br />

will likely gain a stronger position<br />

in the construction market in the<br />

future.<br />

It is also predictable that the methods<br />

of environmental analyses will develop<br />

fast and that their results will enjoy a<br />

growing significance. It is important<br />

to develop objective, soundly based<br />

and well balanced tools enabling us to<br />

comparatively assess the environmental<br />

impacts on engineering choices.<br />

Only such tools can replace emotions,<br />

manipulations and free lobbying,<br />

which very often control these choices<br />

at present. Such tools should be rooted<br />

in official regulations, rather than in<br />

individual judgements. This is the main<br />

reason why the presented assessment<br />

method makes use of “legal thresholds”.<br />

Even if those thresholds are not<br />

perfect yet, they must be endorsed.<br />

The idea behind it is the same as for<br />

referring to the existing databases: it is<br />

better to use them and complain about<br />

their shortcomings than wait until they<br />

improve.<br />

References<br />

[1] Daniel RA. Environmental considerations<br />

to structural material selection for a bridge.<br />

Proceedings of the Bridge Engineering Conference.<br />

COBRAE, Rotterdam, March 2003.<br />

[2] Daniel RA. Construction material for a bridge<br />

with regard to the environment. Bautechnik<br />

2003; 80(1): 32–42.<br />

[3] Daniel RA. Ecological analysis of material<br />

selection for a bridge. Proceedings of<br />

33rd IABSE Symposium. IABSE, Bangkok,<br />

September 2009.<br />

[4] Daniel RA, Brekoo A, Mulder AJ. New<br />

materials for an old navigation lock. Land Water<br />

2001; 41(11): 36–38 (in Dutch).<br />

[5] Schmmelpfeng L, Lück P. Ökologische<br />

Produktgestaltung, Stoffstromanalysen und<br />

Ökobilanzen. Springer-Verlag: Berlin–Heidelberg,<br />

1999.<br />

[6] Elferink H. Energy analysis also useful for<br />

product improving. Energie- en Milieuspectrum<br />

1998; 11: 22–25 (in Dutch).<br />

[7] Anderson JE, Silman R. A life cycle inventory<br />

of structural engineering design strategies<br />

for greenhouse gas reduction. Struct. Eng. Int.<br />

2009; 19(3): 283–288.<br />

[8] Pré Consultants. Environmental Comparison<br />

of Harbour Approach Structures. Research<br />

report 2092. Amersfoort, August 1994 (in Dutch)<br />

(unpublished).<br />

[9] Intron Institute. Energy Analysis for the<br />

Motorway 16/13 Adaptation Options. Research<br />

report M715240/R980548. Houten, November<br />

1994 (in Dutch) (unpublished).<br />

[10] Mahadvi A, Ries R. Towards computational<br />

eco-analysis of building designs. Comput. Struct.<br />

1998; 67: 375–387.<br />

[11] Ministry of Housing, Urban planning<br />

and the environment. The Eco-indicator 99—<br />

Manual for Designers. The Hague, October 2000<br />

(in Dutch).<br />

[12] Hegger M, Fuchs M, Zeumer M. Integration<br />

vergleichender Nachhaltigkeitskennwerte von<br />

Baumaterialien. TU Darmstadt/Deutsche<br />

Bundesstiffung Umwelt, November 2005.<br />

[13] Sittig M. World-wide Limits for Toxic and<br />

Hazardous Chemicals in Air, Water and Soil.<br />

Noyes Publications: Park Ridge, NJ, 1994.<br />

[14] Schlaich M, Bleicher A. Carbon fibre<br />

stress-ribbon bridge. Proceedings of the<br />

COBRAE Conference: Benefits of Composites<br />

in Civil Engineering. COBRAE, Stuttgart,<br />

March 2007.<br />

[15] Daniel RA. Search of Optimal Shapes<br />

for Composite Bridges. Proceedings of the<br />

COBRAE Conference: Benefits of Composites<br />

in Civil Engineering. COBRAE, Stuttgart,<br />

March 2007.<br />

[16] Daniel RA. Shaping composite bridges<br />

for traffic and the environment. Proceedings of<br />

the 4th Int. Conference on Bridge Maintenance,<br />

Safety and Management. IABMAS, Seoul, July<br />

2008. CRC Press, Taylor & Francis Group:<br />

London–New York–Leiden, 2008.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 391


Experimental Assessment of Bond Behaviour<br />

of Fibre-Reinforced Polymers on Brick Masonry<br />

Enrico Garbin, Dr; Matteo Panizza, Dr; Maria Valluzzi, Prof. Dr; University of Padova, Dept. of Structural and Transportation<br />

Engineering, Padova, Italy. Contact: garbin@dic.unipd.it<br />

Abstract<br />

Existing masonry structures represent a significant amount of the architectural<br />

heritage. Many of these buildings are vulnerable to earthquakes. Consequently,<br />

they need structural improvements in order to meet the seismic requirements of<br />

recent building guidelines. In the last decade, there has been a growing interest in<br />

the application of Externally Bonded-Fibre Reinforced Polymers (EB-FRP) as<br />

strengthening and repair materials because of their high-performance mechanical<br />

characteristics, feasibility of application in civil structures, resistance to chemical<br />

attacks and other potentials. Brick masonry components are the most suitable<br />

substrates susceptible to improvements because of their more regular surface<br />

in comparison with stonework or rubble masonry. The bond behaviour of FRP,<br />

applied on a masonry substrate, is a critical issue for the effectiveness of the<br />

technique. In this paper, the results of an experimental assessment of the local<br />

behaviour of EB-FRP applied on clay bricks are presented. Experimental failure<br />

load results were compared with predictive bond strength models proposed in<br />

literature for concrete substrates. On the basis of measured strengths and local<br />

deformations, interface fracture energies were calibrated and an analytical function<br />

was proposed as bond stress-slip law. Finally, a bilinear law was calibrated<br />

for practical design applications.<br />

Keywords: FRP; glass; carbon; bond; masonry; brick.<br />

Introduction<br />

The application of Externally Bonded-<br />

Fibre Reinforced Polymers (EB-FRP)<br />

is a developing technique for the<br />

strengthening and repair of masonry<br />

structures. The bond behaviour between<br />

these products and the substrate is a<br />

crucial aspect that needs to be clarified,<br />

as it strongly influences the effectiveness<br />

of the intervention. In the last<br />

decade, the bonding of composite laminates<br />

on concrete substrates has been<br />

extensively investigated and characterized<br />

using different test set-ups. The<br />

most commonly used are the single-lap<br />

shear test, 1–4 double-lap pull–pull shear<br />

test, 5,6 double-lap push–pull shear test 7<br />

and beam-type test. 8 On the other<br />

hand, few investigations concerning<br />

the debonding from different masonry<br />

and masonry units are available. For<br />

instance refer to investigations on<br />

bond to natural stones, 9 comparison of<br />

natural stones and clay bricks, 10 testing<br />

the bond on solid clay bricks, 11,12<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: February 21, <strong>2010</strong><br />

Paper accepted: July 25, <strong>2010</strong><br />

usage of hollow clay blocks, 7 testing<br />

the bond on brick masonry prisms 13,14<br />

and the research 15 where historic clay<br />

bricks were used as substrate. All of<br />

the aforementioned studies adopted<br />

either the single-lap shear test or the<br />

double-lap push–pull shear test; the<br />

latter has also been referred to in literature<br />

as the double-shear push test<br />

or near-end supported double-shear<br />

test. 16 It involves applying tensile loads<br />

to two reinforcement strips symmetrically<br />

connected to the substrate mainly<br />

to create shear stresses at the interface<br />

while the brittle substrate is subjected<br />

to compressive stresses. The double-lap<br />

push–pull shear test set-up is conventionally<br />

based on the assumption that<br />

the applied load is equally distributed<br />

to the two strips. This test set-up is<br />

quite simple and suitable for commonly<br />

available universal testing machines.<br />

Several predictive models have been<br />

developed to evaluate the debonding<br />

load of composite-to-concrete<br />

joints. Reviews of available strength or<br />

bond-slip models were given by many<br />

researchers 17-20 . The models of Tanaka,<br />

Hiroyuki and Wu, Maeda, Khalifa (as<br />

The models of Tanaka, Hiroyuki and<br />

Wu, Maeda, Khalifa (as reported in<br />

Ref. [17]), Yang, Sato, Iso (as reported<br />

in Ref. [18]), express failure load as the<br />

product of an area and a nominal average<br />

tangential stress. The models of<br />

Izumo, 18 Neubauer and Rostàsy, 17 Chen<br />

and Teng 17 present other expressions<br />

for the failure load. Finally, 11 models<br />

provide an estimation of the fracture<br />

energy value that is correlated with<br />

the failure load. In particular, the models<br />

of Monti (as reported in Ref. [18]),<br />

Lu et al. 18 , herein labelled as Lu Bilinear,<br />

Brosens and Van Gemert 19 and Italian<br />

Research Council Guidelines 21 use a<br />

bilinear bond-slip law; the models of<br />

Nakaba et al. 6 and Savoia et al. 22 adopt<br />

a Popovic’s curve as bond-slip law; the<br />

models of Neubauer and Rostàsy, 17 Dai<br />

& Ueda, 23 Dai et al. 24 and Lu et al. 25<br />

(precise and simplified) resort to other<br />

types of bond-slip function.<br />

Several types of bond-slip laws have<br />

been used in the literature to describe<br />

the behaviour of externally bonded<br />

FRP, namely: (a) a cut-off function<br />

(Neubauer and Rostàsy, as reported<br />

in Ref. [17]); (b) a bilinear function,<br />

presented by some guidelines 26,21 and<br />

by researchers; 18,19,25 (c) a rigid function<br />

with linear softening behaviour; 17<br />

(d) a single function, as the Popovics<br />

curve 22,6 or an exponential curve; 24<br />

(e) two different non-linear functions<br />

for ascending and descending<br />

branches. 25,23 Therefore, it is generally<br />

assumed that bond behaviour of<br />

composite laminates exhibits softening,<br />

with an ascending branch followed<br />

by a descending one, and presents no<br />

residual stress for large slip.<br />

In the present work, the double-lap<br />

push–pull shear procedure was adopted.<br />

Results of five samples for highstrength<br />

carbon reinforcement and five<br />

samples for alkali-resistant glass reinforcement<br />

are presented and discussed.<br />

The predictions of 21 bond strength<br />

models, available in the reported literature<br />

for concrete as the parent material,<br />

were compared with the experimental<br />

debonding loads. The fracture energy<br />

of the composite-to-clay interface was<br />

evaluated using experimental failure<br />

loads. Furthermore, a simplified bondslip<br />

law was proposed on the basis of<br />

the data obtained from the load and<br />

strain recorded during the tests. Finally,<br />

a bilinear function was also calibrated.<br />

392 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Experimental Tests<br />

Solid clay bricks, with nominal dimensions<br />

of 250 × 120 × 55 mm, were used<br />

as the substrate. High-strength Carbon<br />

Fibres Polymer (CFRP) on five specimens<br />

and alkali-resistant Glass Fibres<br />

Polymer (GFRP) on five other specimens<br />

were applied as reinforcement<br />

using a commercially available wet<br />

lay-up system. The mechanical properties<br />

of the bricks used are listed in<br />

Table 1 and the reinforcement system<br />

properties, obtained from manufacturers’<br />

datasheets, are listed in Table 2.<br />

Each specimen was made by a single<br />

clay brick with two strips of reinforcement<br />

symmetrically bonded on the<br />

opposite wider surfaces. The thickness<br />

of the epoxy adhesive was approximately<br />

2 mm and the fibres were placed<br />

as accurately as possible on the centre<br />

line of the polymeric layer. Each strip<br />

was 50 mm wide and bonded to the<br />

brick over a length of L b = 200 mm<br />

(Fig. 1). An unbonded region of 30 mm,<br />

from the loaded end of the brick, was<br />

maintained in order to minimize edge<br />

effects.<br />

The testing machine was a universal<br />

mechanical press (Fig. 2). Each strip<br />

of reinforcement was bonded at the<br />

loaded end to a steel support connected<br />

to the machine. The brick was<br />

connected to the testing apparatus<br />

through a steel frame made by two<br />

steel plates linked by bolts (Fig. 2).<br />

The steel frame and support were connected<br />

to the machine by means of ball<br />

joints, to allow for even loading of the<br />

two composite strips. Samples were<br />

axially loaded in displacement control<br />

at a rate of 0,2 mm/min. The tensile<br />

load was monitored with a 100 kN load<br />

cell. For each specimen, seven strain<br />

gauges were applied to one of the<br />

two reinforcement strips and distributed<br />

as follows: one on the unbonded<br />

region next to the loaded end, and six<br />

on the bonded length. To optimize the<br />

Adhesive Saturant<br />

Characteristic compressive strength >80 N/mm 2<br />

Characteristic direct tensile strength >50 N/mm 2<br />

Maximum tensile strain 2,5%<br />

Tensile modulus of elasticity >3000 N/mm 2<br />

High-strength carbon fibre<br />

Equivalent thickness of one-ply fabric 0,165 N/mm 2<br />

Characteristic direct tensile strength 3430 N/mm 2<br />

Maximum tensile strain 1,5%<br />

Tensile elastic modulus 230 000 N/mm 2<br />

Alkali-resistant glass fibre<br />

Equivalent thickness of one-ply fabric 0,230 N/mm 2<br />

Characteristic direct tensile strength 1700 N/mm 2<br />

Maximum tensile strain 2,8%<br />

Tensile modulus of elasticity 65 000 N/mm 2<br />

Table 2: Properties of reinforcement components<br />

Reinforcement strips<br />

Steel plates<br />

Clay brick<br />

Clay brick<br />

50 mm<br />

Reinforcement<br />

Unbonded zone<br />

200 mm 30 mm<br />

Load direction<br />

Fig. 1: Geometry of specimens<br />

Mean cubic<br />

compressive strength<br />

Mean direct tensile<br />

strength<br />

Mean splitting tensile<br />

strength<br />

Mean flexural tensile<br />

strength<br />

Secant modulus<br />

of elasticity<br />

50,94 N/mm 2<br />

2,37 N/mm 2<br />

3,99 N/mm 2<br />

5,46 N/mm 2<br />

16 100 N/mm 2<br />

Table 1: Mechanical properties of clay<br />

bricks<br />

Fig. 2: Test machine (left) and a specimen ready for testing (right)<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 393


–15 mm<br />

Loaded<br />

end<br />

SG1<br />

Fig. 4: Composite and brick surfaces after test<br />

SG2 SG3 SG4 SG5 SG6 SG7<br />

20 mm<br />

40 mm<br />

65 mm<br />

Clay brick<br />

Fig. 3: Distribution of the strain gauges<br />

95 mm<br />

130 mm<br />

170 mm<br />

Free end<br />

0 200 mm<br />

x axis<br />

and stress were uniformly distributed<br />

along the cross section and that it<br />

was possible to refer to the mechanical<br />

properties of one-ply dry fabric.<br />

This approach is accepted by many<br />

authors 17 and included in some guidelines<br />

21,26 . Therefore, it was possible to<br />

calculate the nominal tensile stress<br />

for each sample as the load divided<br />

by the cross-sectional area, A f =b' f·t f<br />

(where t f is the equivalent thickness of<br />

fibres) and also to calculate the modulus<br />

of elasticity as the linear best fit<br />

of the ratio between the nominal tensile<br />

stress and the experimental strain<br />

data recorded in the unbonded region,<br />

within 10 and 40% of the ultimate load.<br />

The average experimental modulus of<br />

elasticity was found to be higher than<br />

the manufactures’ values (22% for carbon<br />

reinforcement and 25% for glass<br />

reinforcement). Their large dispersion<br />

may be due to the wet lay-up application<br />

of the composites and to minor<br />

misalignments of the fibres. The average<br />

tensile stress reached by the reinforcement<br />

at the debonding load was<br />

64% of the maximum tensile strength<br />

of the carbon fibres and 68% of the<br />

glass fibres.<br />

As the reinforcement axial stiffness<br />

per unit width, E f·t f , was calculated for<br />

each specimen, failure loads per unit<br />

width, P u /2b' f , were tabulated against<br />

the axial stiffness. Trend lines were<br />

fitted, with respect to all data or to<br />

each set (carbon and glass fibres). The<br />

number of instruments and to monitor<br />

the whole bonded region, the strain<br />

gauges were closely spaced near the<br />

loaded end (Fig. 3).<br />

Experimental Test Results<br />

and Prediction of Strength<br />

At the end of the test, all the specimens<br />

showed complete detachment<br />

of the reinforcement from the support.<br />

Failure involved the brick surface<br />

(Fig. 4), where curved cracks<br />

and detachment of clay pieces were<br />

observed. Failure loads, P u , are listed<br />

in Table 3 for carbon (ShC) and glass<br />

(ShG) fibre reinforcements. Specimens<br />

strengthened with CFRP showed 36%<br />

higher failure loads than GFRP specimens.<br />

Table 3 also reports composite<br />

modulus of elasticity values, E f ,<br />

nominal tensile stresses at debonding,<br />

s u , and failure loads per unit width,<br />

P u /2b' f (where b' f is the single strip<br />

width). To compute the modulus of<br />

elasticity, it was assumed that the strain<br />

Specimen E f (N/mm 2 ) P u (N) P u /2b' f (N/mm) r u (N/mm 2 )<br />

High-strength carbon fibre<br />

ShC1 164 419 31 884 318,8 1932<br />

ShC2 336 439 34 233 342,3 2075<br />

ShC3 284 991 35 325 353,3 2141<br />

ShC4 277 511 39 210 392,1 2376<br />

ShC5 338 456 40 301 403 2442<br />

Mean value 280 363 36 191 361,9 2193<br />

Stand. dev. 70 696 3505<br />

COV (%) 25,2 9,7<br />

Alkali-resistant glass fibre<br />

ShG1 50 934 23 380 233,8 1017<br />

ShG2 87 014 27 940 279,4 1215<br />

ShG3 80 545 27 300 273 1187<br />

ShG4 102 598 26 400 264 1148<br />

ShG5 84 842 28 360 283,6 1233<br />

Mean value 81 035 26 676 266,9 1160<br />

Stand. dev. 18 817 1985<br />

COV (%) 23,2 7,4<br />

Table 3: Experimental results for carbon and glass fibre reinforcements<br />

394 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Set of data c 1 c 2 R 2<br />

All data 12,876 0,310 0,876<br />

Glass data only 27,197 0,234 0,622<br />

Carbon data only 35,063 0,218 0,439<br />

All data (square root) 1,759 0,5 n.a.<br />

Glass data (square root) 1,953 0,5 n.a.<br />

Carbon data (square root) 1,681 0,5 n.a.<br />

Table 4: Regression constants of load versus axial stiffness trend lines (both per unit width)<br />

expression adopted for the trend lines<br />

is given in Eq. (1):<br />

P <br />

u<br />

c2<br />

c1<br />

Eftf<br />

<br />

2b'<br />

<br />

= ( ) (1)<br />

f<br />

where c 1 and c 2 are regression constants<br />

(values reported in Table 4). It<br />

can be observed that the fit of all data<br />

shows a better correlation than the fit<br />

of each single set. Moreover, adopting<br />

the relationship between the failure<br />

load and the square root (c 2 equals<br />

0,5) of the axial stiffness per unit width<br />

[Eq. (5)], from the literature 24,27,28 and<br />

Failure load per unit width (N/mm)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

In(y) = 0,21758 In(x) + 3,5572<br />

In(y) = 0,23276 In(x) + 3,30118<br />

In(y) = 0,310003 In(x) + 2,55545<br />

100<br />

Carbon reinforcement<br />

Glass reinforcement<br />

0<br />

0 20 × 10 3 40 × 10 3 60 × 10 3 80 × 10 3<br />

Longitudinal axial stiffness (N/mm)<br />

Fig. 5: Failure loads per unit width versus reinforcement axial stiffness per unit width:<br />

experimental data and trend lines<br />

Failure load per unit width (N/mm)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0<br />

from guidelines 21 , it was possible to<br />

reduce the number of free parameters<br />

in Eq. (1). In this case, the regression<br />

coefficient for the GFRP was slightly<br />

higher than that for CFRP (around<br />

16%) and this could be significant for<br />

the fracture energy evaluation [see<br />

Eq. (6)]. Table 4 gives the values of<br />

carbon data set, glass data set and all<br />

data set, whereas Figs. 5 and 6 compare<br />

trend lines with experimental data.<br />

Experimental results were compared<br />

to estimations of strength given by<br />

the different models presented in the<br />

Introduction section (see Table 5 and<br />

Carbon reinforcement<br />

Glass reinforcement<br />

20 × 10 3 40 × 10 3 60 × 10 3 80 × 10 3<br />

Longitudinal axial stiffness (N/mm)<br />

Fig. 6: Failure loads per unit width versus reinforcement axial stiffness per unit width:<br />

experimental data and trend lines based on the axial stiffness square root<br />

Fig. 7). All the predictions, except<br />

for that of Sato 18 and Izumo 18 in the<br />

in the case of carbon reinforcement,<br />

underestimated the average experimental<br />

failure load. Moreover, all formulations<br />

provided a prediction closer<br />

to test results in the case of CFRP<br />

than in the case of GFRP, except the<br />

models of Tanaka 17 and Hiroyuki 17 .<br />

However, results showed large differences<br />

from model to model. They<br />

varied between 44 and 154% of average<br />

experimental failure load for<br />

carbon reinforcement and between<br />

43 and 85% for glass reinforcement.<br />

Fracture Energy Calibration<br />

The interface mode II fracture energy,<br />

G f , is defined by Eq. (2) as a definite<br />

integral of the tangential stress, t,<br />

expressed as a function of the mutual<br />

slip between the composite and the<br />

substrate, s:<br />

<br />

f<br />

= ( )<br />

G s d s<br />

0<br />

(2)<br />

One of the first analytical models<br />

of the composite-to-concrete bond<br />

strength was derived by Täljsten 29 ,<br />

starting both from a linear approach<br />

based on the beam theory and from a<br />

non-linear approach related to fracture<br />

mechanics. For commonly used epoxy<br />

adhesives, a simplified formulation 29<br />

was obtained [Eq. (3)]:<br />

P<br />

u<br />

= b<br />

f<br />

2EtG<br />

f f f<br />

Et<br />

f f<br />

; <br />

T<br />

= (3)<br />

1+<br />

<br />

Et<br />

T<br />

c c<br />

where b f is the reinforcement width, G f<br />

is the interface fracture energy, a T is<br />

a constant value and E c·t c is the axial<br />

stiffness per unit width of the concrete<br />

substrate.<br />

Yuan (as reported in Ref. [17]) proposed<br />

a modified constant value [Eq. (4)] that<br />

takes into account the width (b f and b c )<br />

ratio of bonded materials:<br />

P<br />

u<br />

= b<br />

f<br />

2EtG<br />

f f f<br />

bEt<br />

f f f<br />

; <br />

W<br />

= (4)<br />

1+<br />

<br />

bEt<br />

W<br />

c<br />

c c<br />

In most cases, the constant value a T , or<br />

a W , has a slight or negligible influence<br />

on the calculation. Several authors 22,24<br />

report the formulation in Eq. (5) without<br />

introducing any constant:<br />

Pu = bf 2 EftfG<br />

(5)<br />

f<br />

By applying Eqs. (3)–(5) to the experimental<br />

data of the present work, it was<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 395


Model<br />

P u /2b' f<br />

(N/mm)<br />

CFRP<br />

Error<br />

(%)<br />

P u /2b' f<br />

(N/mm)<br />

GFRP<br />

Error<br />

(%)<br />

Tanaka 17 166 −54 166 −37,6<br />

Hiroyuki and Wu 17 158 −56,2 158 −40,6<br />

Maeda 17 254 −29,7 174 −34,9<br />

Khalifa 17 248 −31,4 170 −36,2<br />

Yang 18 192 −47 143 −46,3<br />

Sato 18 415 +14,6 116 −56,5<br />

Iso 18 271 −25,1 161 −39,5<br />

Izumo 18 557 +53,9 224 −15,9<br />

Neubauer and Rostàsy 17 283 −21,8 179 −32,7<br />

Chen and Teng 17 245 −32,2 156 −41,6<br />

Monti et al. 18 321 −11,3 204 −23,6<br />

Lu et al. Bilinear 25 220 −39,3 139 −47,7<br />

Brosens and<br />

Van Gemert 19 359 −0,9 228 −14,7<br />

CNR-DT 200 21 263 −27,4 167 −37,5<br />

Nakaba et al. 6 350 −3,3 222 −16,7<br />

Savoia et al. 23 328 −9,4 208 −22<br />

Neubauer and Rostàsy 18 266 −26,4 169 −36,6<br />

Dai and Ueda (1) 23 326 −9,9 207 −22,5<br />

Dai and Ueda (2) 24 322 −11 202 −24,2<br />

Lu et al. Precise 25 220 −39,3 139 −47,7<br />

Lu et al. Simplified 25 220 −39,3 139 −47,7<br />

Mean experimental 362 — 267 —<br />

Table 5: Predictions of failure load<br />

Predicted vs Exp. load ratio<br />

160%<br />

140%<br />

120%<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

0%<br />

Tanaka<br />

HiroyukiI & Wu<br />

Maeda<br />

Khalifa<br />

Yang<br />

Sato<br />

Iso<br />

Izumo<br />

Neubrauer & Rostásy (1)<br />

Chen & Teng<br />

found that taking the parameters a T or<br />

a W into consideration only leads to a<br />

difference of less than 2%. It is worth<br />

noting that Eq. (5), demonstrated in<br />

some cases, 24,27 can be assumed in every<br />

case of regular interface law as pointed<br />

out by Savoia et al. 28 The formulation<br />

of Eq. (5), derived for concrete<br />

substrates, has been considered valid<br />

also for the clay substrate adopted in<br />

the present work, since both are quasibrittle<br />

substrates. Accordingly, it was<br />

possible to calibrate the fracture energy<br />

G f through Eq. (5), using mean values<br />

Monti et al.<br />

Lu et al. Bilinear<br />

Brosens & Van Gemert<br />

Carbon reinforcement<br />

Glass reinforcement<br />

CNR DT–200/2004<br />

Nakaba<br />

Savoia<br />

Neubauer & Rostásy (2)<br />

Dai & Ueda (1)<br />

Dai & Ueda (2)<br />

Lu et al. Precise<br />

Lu et al. Simplified<br />

Fig. 7: Ratio of predicted failure loads versus average experimental value for carbon and<br />

glass reinforcement<br />

of failure load and elastic modulus, and<br />

the results are given in Table 6. The estimated<br />

value for glass reinforcement is<br />

approximately 35% higher than that<br />

for carbon reinforcement.<br />

Moreover, the fitting parameter c 1<br />

given in Table 4, when c 2 is 0,5 (square<br />

root based fit), allowed evaluation of<br />

G f as shown in Eq. (6). The results in<br />

Table 6 presented no significant difference<br />

from the values obtained by<br />

means of Eq. (5) for carbon and glass<br />

fibres, while using all data the value<br />

provided by Eq. (5) was 11% higher<br />

than the value provided by Eq. (6):<br />

2<br />

2EtG f f f= c1<br />

Et<br />

f f⇒ Gf= 0, 5c<br />

(6)<br />

1<br />

Calibration of a Bond-Slip Law<br />

Calibration of the bond-slip law on the<br />

basis of the experimental results was<br />

performed by adopting a combined<br />

approach where the tangential stress<br />

and interface slip points (t –s) were<br />

obtained from strain gauge recordings<br />

and the fracture energy value, G f , was<br />

calculated from failure loads by means<br />

of Eq. (5). The fracture energy represents<br />

an analytical restraint for the<br />

bond-slip function [Eq. (2)] and allows<br />

reduction of the number of free parameters<br />

involved in the calibration process.<br />

Equations (7)–(9), which briefly report<br />

the main relations among reinforcement<br />

strain e, interface tangential<br />

stress t and slip s, were obtained from<br />

simple equilibrium and compatibility<br />

considerations, disregarding the slip<br />

component due to the substrate, which<br />

is generally stiffer than the composite.<br />

The notation x indicates the coordinate<br />

along the central axis of the<br />

bonded region.<br />

d<br />

x<br />

dx<br />

( ) = ( )<br />

1<br />

x<br />

E t<br />

f f<br />

dsx<br />

( x)=<br />

dx<br />

2<br />

d sx<br />

dx<br />

x<br />

( ) ( )= ( )<br />

( ) ( )=<br />

f f<br />

<br />

sx xdx<br />

0<br />

(7)<br />

(8)<br />

1 x 0<br />

(9)<br />

E t<br />

To calculate the tangential stress and slip<br />

values from the corresponding strain<br />

recorded in discrete positions along the<br />

reinforcement, Eqs. (7)–(9) were modified<br />

and input into the discrete formulas<br />

given in Eqs. (10) and (11), as was done<br />

396 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Reinforcement type G f from Eq. (5) (N/mm) G f from S.R. fit (N/mm)<br />

Carbon fibres 1,42 1,41<br />

Glass fibres 1,91 1,91<br />

All data 1,72 1,55<br />

Table 6: Evaluation of fracture energy<br />

in Ref. [30]. This allows for the manipulation<br />

of non-uniformly spaced data:<br />

1<br />

<br />

i<br />

= ( xi)=<br />

E t<br />

2<br />

f f<br />

i<br />

i<br />

i<br />

<br />

+ 1 1<br />

i<br />

+<br />

<br />

x x x x <br />

<br />

i+<br />

1<br />

i<br />

i<br />

i1<br />

(10)<br />

1<br />

si = s( xi)= si+ 1+ ( i<br />

+<br />

i+ 1) ( xi+<br />

1<br />

xi)<br />

2 (11)<br />

where the notation i − n indicates the<br />

strain gauge position. The x axis is<br />

oriented such that i increases from<br />

the loaded end (x = 0) to the free end<br />

(x = 200 mm).<br />

As explained in the Introduction section,<br />

it is assumed that the bond-slip law<br />

should show an ascending segment and<br />

a softening behaviour. Instead of using<br />

two different mathematical expressions<br />

for the ascending and the descending<br />

branches, a single function was selected<br />

on purpose. Although there could be a<br />

slight loss of adherence to experimental<br />

data, the single function reduces the<br />

required parameters, making the fitting<br />

easier. The proposed law, easy to integrate<br />

and derive, is given in Eq. (12):<br />

Bs<br />

( s)= Ase (12)<br />

where A and B are regression constants,<br />

t is the interface tangential stress and<br />

s is the slip. By applying the calibrated<br />

fracture energy value, it was possible<br />

to fit a function that depended on just<br />

one parameter, as shown in Eq. (13):<br />

After the optimization of the UniPd<br />

curves for carbon and glass reinforcements,<br />

it was also possible to calibrate<br />

a bilinear bond-slip law, which is commonly<br />

adopted by some guidelines<br />

(FIB Bulletin 26 ; CNR-DT 200 21 ). The<br />

analytical form of the bilinear law is<br />

reported in Eq. (15):<br />

( ) <<br />

<br />

max<br />

ss0 0 s s 0<br />

<br />

( s)=<br />

<br />

max (( s s) ( s s0<br />

)) s s<<br />

s<br />

<br />

<br />

0 s s f<br />

f f 0 f<br />

(15)<br />

where s f is the ultimate slip related to<br />

null tangential bond stress.<br />

Bond strees (N/mm 2 )<br />

8,0<br />

6,0<br />

4,0<br />

2,0<br />

0,0<br />

0,000 0,200 0,400<br />

Slip (mm)<br />

Fig. 8: Calibrated bond-slip laws for CFRP reinforcement<br />

As the bilinear function depends on<br />

more parameters than the UniPd<br />

curve, the maximum tangential stress<br />

value, obtained from the fitted UniPd<br />

curve, and the calibrated fracture<br />

energy were imposed during the optimization<br />

process. Figures 8 and 9 show<br />

the optimized curves and the experimental<br />

stress-slip data. It can be noted<br />

that carbon reinforcement interface<br />

seems to be slightly stiffer than the<br />

glass reinforcement.<br />

Tables 7 and 8 report the significant<br />

values (fracture energy, peak tangential<br />

stress with related slip and ultimate<br />

slip) calculated by fitting the<br />

experimental data. They were compared<br />

with the values estimated using<br />

the 11 models reported in the literature<br />

and based on the fracture energy<br />

prediction. Estimated values varied<br />

across a wide range. Most models did<br />

not provide any difference for CFRP<br />

and GFRP reinforcements. Compared<br />

to the experimental results, they tend<br />

to underestimate the fracture energy<br />

(from 2 to 72%) and often do not<br />

CFRP<br />

UniPD curve<br />

Bilinear<br />

0,600 0,800<br />

∞<br />

A<br />

2<br />

∫ τds<br />

= ⇒ A=<br />

B Gf<br />

;<br />

2<br />

B<br />

0<br />

2<br />

− Bs<br />

τ s B G s e<br />

()= ⋅ ⋅<br />

f<br />

(13)<br />

The bond-slip law [Eq. (12)], herein<br />

labelled UniPd curve, can be rewritten<br />

in a normalized form as generally used<br />

in guidelines [Eq. (14)]:<br />

( s)=<br />

<br />

max<br />

<br />

s <br />

s 1<br />

s<br />

<br />

0 <br />

<br />

s <br />

e (14)<br />

0<br />

where s 0 = 1/B and t max = t (s 0 ) are the<br />

coordinates of the maximum tangential<br />

stress point.<br />

Bond strees (N/mm 2 )<br />

8,0<br />

6,0<br />

4,0<br />

2,0<br />

0,0<br />

0,000 0,200 0,400<br />

Slip (mm)<br />

Fig. 9: Calibrated bond-slip laws for GFRP reinforcement<br />

CFRP<br />

UniPD curve<br />

Bilinear<br />

0,600 0,800<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 397


Curve G f (N/mm) s max (N/mm 2 ) s 0 (mm) s f (mm)<br />

UniPd fitting 1,42 7,22 0,072 —<br />

Bilinear fitting 1,42 7,22 0,034 0,392<br />

Monti et al. 18 1,11 5,37 0,046 0,415<br />

Lu et al. Bilinear 25 0,52 3,73 0,048 0,28<br />

Brosens and<br />

Van Gemert 19 1,39 2,71 0,012 1,025<br />

CNR-DT 200 21 0,75 7,46 0,056 0,2<br />

Nakaba et al. 6 1,32 7,08 0,065 —<br />

Savoia et al. 23 1,16 7,08 0,051 —<br />

Neubauer and<br />

Rostàsy 18 0,77 5,69 0,27 —<br />

Dai and Ueda (1) 23 1,15 8,58 0,103 —<br />

Dai and Ueda (2) 24 1,12 6,41 0,061 —<br />

Lu et al. Precise 25 0,52 3,73 0,054 —<br />

Lu et al. Simplified 25 0,52 3,73 0,048 —<br />

Table 7: Significant values for local bond of CFRP<br />

Curve G f (N/mm) s max (N/mm 2 ) s 0 (mm) s f (mm)<br />

UniPd fitting 1,91 6,33 0,111 —<br />

Bilinear fitting 1,91 6,33 0,048 0,603<br />

Monti et al. 18 1,11 5,37 0,046 0,415<br />

Lu et al. Bilinear 25 0,52 3,73 0,048 0,28<br />

Brosens and<br />

1,39 2,71 0,012 1,025<br />

Van Gemert 19<br />

CNR-DT 200 21 0,75 7,46 0,056 0,2<br />

Nakaba et al. 6 1,32 7,08 0,065 —<br />

Savoia et al. 23 1,16 7,08 0,051 —<br />

Neubauer and<br />

0,77 5,69 0,27 —<br />

Rostàsy 18<br />

Dai and Ueda (1) 23 1,15 7,1 0,107 —<br />

Dai and Ueda (2) 24 1,1 5,69 0,067 —<br />

Lu et al. Precise 25 0,52 3,73 0,054 —<br />

Lu et al. Simplified 25 0,52 3,73 0,048 —<br />

Table 8: Significant values for local bond of GFRP<br />

correctly estimate the maximum tangential<br />

stress and ultimate slip.<br />

Conclusion<br />

The bond behaviour of composite-to-clay<br />

brick interface was investigated using<br />

double-lap push–pull shear tests, for<br />

both high-strength carbon (CFRP) and<br />

alkali-resistant glass (GFRP) reinforcements.<br />

Far from being exhaustive, the<br />

experimental work was mainly focused<br />

on setting a procedure to design, perform<br />

and analyse this local phenomenon.<br />

Strength results showed a better performance<br />

of carbon reinforcement than<br />

glass, around 36% higher in the first<br />

case.<br />

Experimental strengths were compared<br />

with those obtained from 21 predictive<br />

models developed for concrete<br />

substrate. All predictions, except two in<br />

the case of CFRP, underestimated the<br />

test results. Models, except two in case<br />

of GFRP, appeared to provide better<br />

estimations for carbon reinforcement.<br />

However, the range of strength predictions<br />

was rather wide (between 44 and<br />

154% of average experimental failure<br />

load for CFRP and between 43 and<br />

85% for GFRP).<br />

From measured failure loads, different<br />

fracture energy values were obtained,<br />

around 35% higher in the case of glass<br />

reinforcement than carbon reinforcement.<br />

A mathematical function, easy to<br />

integrate and derive, was proposed for<br />

a bond-slip law and fitted for both carbon<br />

and glass reinforcement. Finally,<br />

two bilinear functions were also calibrated<br />

for design purposes according<br />

to strengthening guidelines. The optimized<br />

functions seem to show a bond<br />

behaviour for CFRP that is slightly<br />

stiffer than for GFRP.<br />

Further investigations are ongoing<br />

within the framework of the<br />

Rilem Technical Committee 223-<br />

MSC “Masonry Strengthening with<br />

Composite materials”, aimed at deepening<br />

the knowledge on the present<br />

topic and defining specific standardized<br />

procedures for testing the adhesion<br />

of composite materials applied to<br />

masonry.<br />

Acknowledgements<br />

The authors wish to thank BASF CC Italia<br />

of Treviso, Italy, for the technical collaboration<br />

and for supplying fibres and the adhesion<br />

system. The research activity has been<br />

also partially supported by the National<br />

Italian Project ReLUIS. The authors would<br />

like to thank Eng. A. Cartolaro and the staff<br />

of the Laboratory of Material Testing of the<br />

Department of Structural and Transportation<br />

Engineering of the University of Padova,<br />

where tests were carried out.<br />

References<br />

[1] Chajes MJ, Finch WW, Januszka TF, Thomson<br />

TA. Bond and force transfer of composite material<br />

plates bonded to concrete. ACI Struct. J.<br />

1996; 93(2): 295–303.<br />

[2] Täljsten B. Defining anchor lengths of steel<br />

and CFRP plates bonded to concrete. Int. J.<br />

Adhes. Adhes. 1997; 19: 319–327.<br />

[3] Bizindavyi L, Neale KW. Transfer lengths and<br />

bond strengths for composites bonded to concrete.<br />

J. Compos. Const. 1999; 3(4): 153–160.<br />

[4] Mazzotti C, Savoia M, Ferracuti B. A new single-shear<br />

set-up for stable debonding of FRP –<br />

concrete joints. Const. Build. Mater. 2009; 23:<br />

1529–1537.<br />

[5] Lee YJ, Boothby TE, Bakis CE, Nanni A.<br />

Slip modulus of FRP sheets bonded to concrete.<br />

ASCE J. Compos. Constr. 1999; 3(4): 161–167.<br />

[6] Nakaba K, Kanakubo T, Furuta T, Yoshizawa<br />

H. Bond behavior between fiber-reinforced<br />

polymer laminates and concrete. ACI Struct. J.<br />

2001; 98(3): 359–367.<br />

[7] Camli US, Binici B. Strength of carbon<br />

fiber reinforced polymers bonded to concrete<br />

and masonry. Const. Build. Mater. 2007; 21:<br />

1431–1446.<br />

[8] De Lorenzis L, Miller B, Nanni A. Bond of<br />

FRP laminates to concrete. ACI Mater. J. 2001;<br />

98(3): 256–264.<br />

[9] Aiello MA, Sciolti MS. Bond analysis of<br />

masonry structures strengthened with CFRP<br />

sheets. Const. Build. Mater. 2006; 20: 90–100.<br />

398 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


[10] Faella C, Martinelli E, Paciello S, Perri F.<br />

Composite materials for masonry structures:<br />

the adhesion issue. In MuRiCo 3: Proceedings<br />

of the 3rd Nat. Conf. on Mechanics of Masonry<br />

Structures Strengthened with Composite<br />

Materials: Modeling, Testing, Design, Control;<br />

2009 April 1–3, Venice, Italy. Pitagora, Bologna,<br />

Di Tommaso A. (ed.), 2009; 266–273.<br />

[11] Briccoli Bati S, Rovero L, Tonietti U.<br />

Adesione fra blocchi in laterizio e rinforzo in<br />

CFRP. In Proceedings of the National Workshop<br />

on Materiali ed Approcci Innovativi per il<br />

Progetto in Zona Sismica e la Mitigazione della<br />

Vulnerabilità delle Strutture, Faella C, Manfredi<br />

G, Piluso V, Realfonzo R (eds), Salerno, Italy.<br />

Monza: Polimetrica, 2007; 213–220.<br />

[12] Panizza M. FRP Strengthening or Masonry<br />

Arches: Analysis of Local Mechanisms and<br />

Global Behaviour. Ph.D. [dissertation]. Padova<br />

(Italy): University of Padova; <strong>2010</strong>.<br />

[13] Basilio I. Strengthening of Arched Masonry<br />

Structures with Composite Materials. Ph.D. [dissertation].<br />

Guimarães (Portugal): University of<br />

Minho; 2007.<br />

[14] Subramanian K, Focacci F, Carloni C.<br />

An investigation on the interface fracture propagation<br />

between FRP and masonry. In MuRiCo 3:<br />

Proceedings of the 3 rd Nat. Conf. on Mechanics of<br />

masonry structures strengthened with composite<br />

materials: modeling, testing, design, control; 2009<br />

April 1-3, Venice, Italy. Pitagora, Bologna, Di<br />

Tommaso A (ed.), Bologna, 2009; 423–430.<br />

[15] Capozucca R. Experimental FRP/SRPhistoric<br />

masonry delamination. Compos Struct.<br />

<strong>2010</strong>; 92(4): 891-903.<br />

[16] Yao J, Teng JG, Chen JF. Experimental study<br />

on FRP-to-concrete bonded joints. Compos: B<br />

2005; 36: 99–113.<br />

[17] Chen JF, Teng JG. Anchorage strength models<br />

for FRP and steel plates bonded to concrete.<br />

ASCE J. Struct. Eng. 2001; 127(7): 784–791.<br />

[18] Lu XZ, Teng JG, Ye LP, Jiang JJ. Bond-slip<br />

models for FRP sheets-plates bonded to concrete.<br />

Eng. Struct. 2005; 27: 920–937.<br />

[19] Karbhari VM, Niu H, Sikorsky C. Review<br />

and comparison of fracture mechanics-based<br />

bond strength models for FRP-strengthened<br />

structures. J. Reinforced Plastics Compos. 2006;<br />

25(17): 1757–1794.<br />

[20] Ferracuti B, Savoia M, Mazzotti C. Interface<br />

law for FRP—concrete delamination. Composite<br />

structures. 2007; 80(4), 523–531.<br />

[21] CNR-DT 200/2004. Guide for the Design<br />

and Construction of Externally Bonded FRP<br />

Systems for Strengthening Existing Structures.<br />

Italian National Research Council: Rome, 2005.<br />

[22] Savoia M, Ferracuti B, Mazzotti C. Non linear<br />

bond-slip law for FRP-concrete interface. In:<br />

FRPRCS-6: Proceedings of the 6th International<br />

Symposium on Fibre-Reinforced Polymer<br />

Reinforcement for Concrete Structures; 2003 July<br />

8–10; Singapore, Tan KH (ed.). World Scientific:<br />

Singapore, 2003; 1–10.<br />

[23] Dai JG, Ueda T. Local Bond stress Slip relations<br />

for FRP Sheets-Concrete Interfaces. In:<br />

FRPRCS-6: Proceedings of the 6th International<br />

Symposium on Fibre-Reinforced Polymer<br />

Reinforcement for Concrete Structures; 2003 July<br />

8–10; Singapore, Tan KH (ed.), World Scientific:<br />

Singapore, 2003; 143–152.<br />

[24] Dai JG, Ueda T, Sato Y. Development of the<br />

nonlinear bond stress-slip model of fiber reinforced<br />

plastics sheet–concrete interfaces with a<br />

simple method. ASCE J Compos Const. 2005;<br />

9(1): 52–62.<br />

[25] Lu XZ, Teng JG, Ye LP, Jiang JJ. Bondslip<br />

models for FRP sheet/plate-to-concrete<br />

interfaces. In: ACIC 2004: 2nd International<br />

Conference on Advanced Polymer Composites<br />

for Structural Applications in Construction; 2004<br />

April 20–22; Guildford, Great Britain, Hollaway<br />

LC, Chryssanthopoulos MK, Moy SSJ (eds).<br />

Woodhead: Cambridge, 2004; 152–161.<br />

[26] FIB Bulletin 14. Externally Bonded FRP<br />

Reinforcement for RC Structures. Fédération<br />

Internationale du Béton: Lausanne, 2001.<br />

[27] Wu Z, Yuan H, Niu H. Stress transfer and<br />

fracture propagation in different kinds of adhesive<br />

joints. ASCE J. Eng. Mech. 2002; 128(5):<br />

562–573.<br />

[28] Savoia M, Ferracuti B, Mazzotti C. Una<br />

legge di interfaccia non lineare per placcaggi<br />

con lamine in FRP. In: AIMETA’03: Proceedings<br />

of the 16th Congress of Theoretical and Applied<br />

Mechanics; Sep 9–12; Ferrara, Italy, 2003.<br />

[29] Täljsten B. Strengthening of concrete prisms<br />

using the plate bonding technique. Int. J. Fract.<br />

1996; 82: 253–266.<br />

[30] Valluzzi MR, Tinazzi D, Garbin E,<br />

Modena C. FEM modelling of CFRP strips<br />

bond behaviour for bed joints reinforcement<br />

techniques. In: STRUMAS VI: Proceeding of<br />

the 6th International Conference on Computer<br />

Methods in Structural Masonry; Sep 22–24; Rome,<br />

Italy, 2003.<br />

Further Information<br />

http://www.rilem.net/tcDetails.php?tc=223-MSC<br />

http://www.cnr.it/sitocnr/Englishversion/CNR/<br />

Activities/RegulationCertification.html<br />

http://www.reluis.it/<br />

List of Symbols<br />

A Regression constant<br />

A f Cross-section area of a single strip<br />

B Regression constant<br />

b c Width of the concrete<br />

b f Reinforcement width<br />

b' f Width of a single strip<br />

c 1 Regression constant<br />

c 2 Regression constant<br />

e Euler’s Constant<br />

E c Modulus of elasticity of concrete<br />

E f Longitudinal modulus of elasticity of the<br />

fibres<br />

G f Interface fracture energy in mode II<br />

L b Bonded length<br />

P u Failure load<br />

s Interface relative slip<br />

s 0 Interface slip at the maximum tangential<br />

bond stress<br />

s f Ultimate interface slip at null tangential<br />

bond stress<br />

t c Thickness of the concrete<br />

t f Thickness of the fibres<br />

x Longitudinal abscissa along the bonded<br />

length<br />

a T Constant value<br />

a W Constant value<br />

e Reinforcement strain<br />

s u Nominal tensile stresses at debonding<br />

t Tangential bond stress<br />

t max Maximum tangential bond stress<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 399


Bridges with Glass Fibre–Reinforced Polymer Decks:<br />

The Road Bridge in Friedberg, Germany<br />

Jan Knippers, Head of Institute, Universität Stuttgart, ITKE, Stuttgart, Germany; Eberhard Pelke, Head of Dept., Road and Traffic<br />

Administrative Dept. Hessen, Wiesbaden, Germany; Markus Gabler, Research Assistant, Universität Stuttgart, ITKE, Stuttgart,<br />

Germany; Dieter Berger, Development Eng., Road and Traffic Administrative Department Hessen, Wiesbaden, Germany.<br />

Contact: info@itke.uni-stuttgart.de<br />

Abstract<br />

In July 2008, the first road bridge in<br />

Germany using glass fibre-reinforced<br />

polymers (GFRP) was completed in<br />

Friedberg/Hessen. The structure has<br />

a span of 27 m and acts as a flyover<br />

across the federal road B3. The high<br />

durability of the new construction<br />

material and the fast assembly of the<br />

bridge were decisive factors in favour<br />

of GFRP.<br />

During the preceding years, several<br />

lightweight bridges using FRP had<br />

been constructed in the USA, Japan<br />

and also in Europe. Through these<br />

projects, valuable experience was gathered<br />

regarding construction, and the<br />

use and performance of composites<br />

could be demonstrated.<br />

The bridge in Friedberg extended this<br />

experience by taking into account the<br />

composite action between the FRP<br />

deck and the steel girders. It also<br />

followed, consequently, the approach<br />

of durable bridge construction by omitting<br />

any bearings or expansion joints<br />

and making the innovative material<br />

visible to passers-by.<br />

Keywords: glass fibre-reinforced polymers;<br />

composites; road bridge; installation;<br />

material tests; approval.<br />

Introduction<br />

The road and traffic administrative<br />

department in the German state of<br />

Hessen sees itself as a modern mobility<br />

service provider. It consequentially<br />

supports the development of new<br />

building methods for civil engineering<br />

structures, providing sustainability and<br />

reduction of life cycle costs as well as<br />

minimal interference with traffic during<br />

the construction. This includes the<br />

application of new materials in building<br />

methods.<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: February 26, <strong>2010</strong><br />

Paper accepted: July 1, <strong>2010</strong><br />

As the “hub of Germany” due to its<br />

central position, the state of Hessen<br />

has the highest average volume of<br />

traffic in Germany. Durability and<br />

low maintenance of civil engineering<br />

structures are hence prerequisites in<br />

terms of maintaining the economy’s<br />

efficiency and reducing costs of citizens<br />

by keeping traffic congestions to<br />

a minimum. The first approaches were<br />

conducted by the administration by<br />

observing traditional bridge building. 1<br />

Today, project calculations of road<br />

construction include approximate life<br />

cycle costs. In order to implement<br />

the objectives mentioned above, different<br />

new building techniques were<br />

examined and analysed. Lightweight<br />

constructions using fibre-reinforced<br />

polymer (FRP), which is very durable<br />

and fast to assemble, provide a potential<br />

solution. Such a bridge system was<br />

recently developed and realised in the<br />

context of the project “Congestion-<br />

Free Hessen”.<br />

Bridges with FRP Decks<br />

Realised Constructions<br />

The costs of FRP products are higher<br />

than that of conventional materials,<br />

therefore it is sensible to restrict their<br />

use to members that are susceptible to<br />

corrosion. For bridge superstructures,<br />

this means that decks can be built out of<br />

FRP, while the main girders are made<br />

of conventional materials such as steel<br />

and reinforced concrete (RC). This<br />

means not only lower costs but also less<br />

deflection of the superstructure due to<br />

the higher Young’s modulus of steel<br />

and RC. Steel FRP-composites have,<br />

especially in the USA, been extensively<br />

applied since the mid 1990s. The<br />

US Federal Highway Agency presently<br />

counts approximately 70 bridges<br />

built with FRP decks. 2 FRP decks are<br />

used very often in the refurbishment<br />

of existing road bridges. Damaged<br />

decks have been replaced on numerous<br />

RC-composite or steel bridges by<br />

FRP panels, whereas the abutment<br />

and the steel beams were preserved.<br />

The possibility of prefabricating large<br />

lightweight panels and transporting<br />

them to the building site, allows for<br />

extremely rapid refurbishment and<br />

reopening of the bridges. The curing<br />

time necessary for concrete decks is<br />

a fundamental reason for choosing<br />

the application of FRP. Several road<br />

bridges with FRP decks have also<br />

been built in Korea in the last few<br />

years, the 300 m long and 35 m wide<br />

“Noolcha Bridge” in Busan, for example.<br />

3 Positive experience was thereby<br />

consistently gathered—no major damages<br />

due to long term load have been<br />

observed in the structures built so far. 4<br />

Types of FRP Decks<br />

Available FRP decks can generally<br />

be classified into two categories: pultruded<br />

hollow sections and hand layup<br />

sandwich panels (Fig. 1). Pultruded<br />

decks consist of a row of prismatic<br />

bars that are manufactured through an<br />

automatic process. These hollow sections<br />

have a wall thickness that varies<br />

from 5 to 15 mm and an overall dimension<br />

of approximately 200 mm × 400<br />

mm. Owing to the continuous manufacturing<br />

process, the fibre direction<br />

is mostly longitudinal. The panels are<br />

formed by adhesively bonding the bars<br />

together. The bars are connected to the<br />

slab during assembly or previously at<br />

the factory in transportable sizes. The<br />

pultruded deck sections generally have<br />

a span of 2 to 3 m between the main<br />

girders.<br />

On the other hand, sandwich panels<br />

bear loads in two directions with the<br />

same stiffness and are therefore better<br />

suited to carrying concentrated<br />

loads. However, bigger tolerances<br />

occur through the manual manufacturing<br />

process than in pultruded profiles.<br />

Furthermore, research has shown that,<br />

owing to minimal heat transfer in the<br />

sandwich panels, very high temperature<br />

gradients can occur. Lastly, the<br />

intricate construction of connections<br />

and the connecting details are more<br />

challenging, which has led to sandwich<br />

panels not being introduced in Europe<br />

to date.<br />

400 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Fig. 1: Comparison of different systems for FRP bridge decks; (left): pultruded hollow sections;<br />

(right): manually laid-up sandwich panels<br />

Dimensioning and Building<br />

Draft Principles<br />

Unlike the case of previously built<br />

structures, the aim of the design for<br />

the new road bridge in Friedberg,<br />

Germany, was to fully apply the composite<br />

action between the steel girders<br />

and the FRP deck, and to verify this<br />

experimentally as well as through analysis.<br />

For this reason, an adhesive layer<br />

was chosen for the FRP–steel connection:<br />

a joining method so far rarely<br />

used for civil engineering structures,<br />

although the material is appropriate<br />

and enables smooth load transmission<br />

without weakening caused by drilling.<br />

A frame structure was chosen with the<br />

purpose of long life and low upkeep of<br />

the construction, and therefore, track<br />

transition and bridge bearings were<br />

omitted. The structure was designed<br />

according to the ENV 1991–3 5 taking<br />

into account the government regulations<br />

of Germany. Due to lack of<br />

appropriate data, the fatigue-proof<br />

test of FRP could not be carried out<br />

through calculation. It was substituted<br />

by test data from the manufacturer of<br />

the FRP deck, which certified advantageous<br />

fatigue behaviour.<br />

Structural System and Composition<br />

The bridge structure (Fig. 2) has an<br />

integral design with a span of 27 m<br />

over the two-lane federal road B3a<br />

near Friedberg, in central Germany.<br />

The superstructure was made of two<br />

slightly curved and haunched steel<br />

girder with a structural height of<br />

625 to 900 mm. The overall 5 m wide<br />

FRP deck was glued onto both main<br />

girders (Fig. 3). A two-component<br />

epoxy adhesive was used to attach<br />

the FRP to the steel. The deck consisted<br />

of glued hollow sections and<br />

had a structural height of 225 mm. It<br />

was covered by a 45 mm thick layer<br />

of polymer concrete with applied<br />

polymer surfacing, which was sprayed<br />

with silica to provide surface roughness.<br />

The tensile strength of the polymer<br />

concrete was high enough to<br />

be considered in the calculation as<br />

strengthening for the top flange of the<br />

FRP deck. To obtain sufficient loadbearing<br />

safety, such composite action<br />

was necessary for withstanding concentrated<br />

wheel loads. The edge casings<br />

were made with a second layer of<br />

the same FRP sections. These casings<br />

together with the steel railing guarantee<br />

the required safety for the cars on<br />

the bridge. The design velocity of the<br />

road passing over the bridge is V ≤ 50<br />

km/h. The railing itself was connected<br />

to the front faces with anchor plates<br />

that were glued into the hollow sections<br />

of the FRP profiles of the deck<br />

and edge casings. Thus, it is possible<br />

to replace these parts in case of damage.<br />

The front face of the FRP decks<br />

was closed with thin-walled plates in<br />

order to prevent vermin from entering<br />

the hollow sections (Fig. 4).<br />

After assembling, the entire superstructure<br />

was rigidly connected onto<br />

the concrete abutments by grouting<br />

the attached fixing plates with<br />

welded head stud bolts. Making the<br />

FRP parts visible for passers-by was<br />

of utmost importance; therefore the<br />

front faces of the FRP deck remained<br />

uncovered, the hollow parts being<br />

recognisable as shadow gaps. The<br />

frame appearance of the bridge was<br />

underlined by the choice and orientation<br />

of formwork.<br />

625<br />

7000 20000 7000<br />

Fig. 2: FRP bridge in Friedberg (Hessen—Germany); (top): elevation; (bottom): longitudinal<br />

section (Units: mm)<br />

900<br />

8290<br />

Finite Element Method (FEM)—<br />

Calculations and Laminate Analysis<br />

The bridge structure was designed as a<br />

restrained frame with linear structural<br />

members. The composite cross section<br />

of the superstructure was therefore<br />

represented as an ideal steel cross section.<br />

It was already known from available<br />

test results 6 that full composite<br />

action can be assumed for the adhesive<br />

layer and the deck as the cross<br />

section virtually stayed even. With the<br />

actual dimensions of the Friedberg<br />

superstructure, the entire width of the<br />

FRP deck is under compression in<br />

the centre span. Owing to the direction<br />

of the fibre reinforcement, the<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 401


East<br />

1300<br />

+245<br />

+45<br />

2,5%<br />

Steel girder (h = 625 ... 900)<br />

1200 1200<br />

5000<br />

slab mainly acts in uniaxial direction<br />

perpendicular to the bridge span. The<br />

load transfer presumptions were verified<br />

through tests.<br />

Specific values of material properties<br />

are necessary for the calculation of<br />

FRP structures as, on the one hand,<br />

different laminate set-ups are used for<br />

top flanges and webs of the deck and,<br />

on the other hand, the axial stiffness<br />

differs from the values in the lateral<br />

direction. A laminate calculation was<br />

therefore performed as a preparation<br />

to the actual analysis, determining the<br />

existing Young’s modulus and ultimate<br />

strength of the components based on<br />

the background of the respective fibre<br />

architecture (Table 1).<br />

For the input values of actions on<br />

bridges, in addition to ENV 1991–3,<br />

assumptions regarding the following<br />

issues had to be made:<br />

– For the additional ultimate limit<br />

state of creep rupture, 100% of dead<br />

load, 20% of imposed load and 50%<br />

of temperature load were taken into<br />

account and opposed to a reduced<br />

material strength.<br />

7 wearing course<br />

45 polymer concrete<br />

+155<br />

– 45<br />

Fig. 3: Standard cross section of composite superstructure (Units: mm)<br />

Fig. 4: Front face of superstructure (photo<br />

credit: Wilfried Dechau)<br />

– Rupture of an adhesively bonded<br />

joint was taken into account as an<br />

accidental action.<br />

– Analysis of fatigue loads was omitted<br />

due to existing verification<br />

through tests of the fatigue strength<br />

by the manufacturer.<br />

– Climatic temperature changes on<br />

steel–FRP composite superstructure<br />

were assumed to be T e = −20 to 41 K<br />

(overall cross section) and ΔT M =<br />

−18/+25 K (temperature moment).<br />

On the materials side, there are still gaps<br />

in building codes for FRP structures.<br />

In Germany, the recommendations of<br />

BUV (“BUV-Empfehlungen” 7 ) provide<br />

valuable design criteria. Structural<br />

Design of Polymer Composites—<br />

EUROCOMP Design Code and<br />

Handbook 8 is a helpful handbook for<br />

engineers. Furthermore, additional<br />

assumptions and alterations had to<br />

be considered for the analysis of the<br />

Friedberg Bridge:<br />

– superimposition of stress resultants<br />

according to the Puck/Knaust relationship<br />

of interaction;<br />

– relation of interaction at adhesively<br />

bonded connections;<br />

402 Scientific Paper Structural Engineering International 4/<strong>2010</strong><br />

1300<br />

4%<br />

1000<br />

FRP-deck<br />

West<br />

N/mm 2 E xx E yy G xy f x,u,k f y,u,k<br />

Flange 28,000 19,000 5,000 250 175<br />

Outer web 17,000 24,000 4,500 160 125<br />

Inner web 17,000 26,000 3,000 210 195<br />

x: pultrusion direction y: Lateral to pultrusion direction<br />

Table 1: Properties of utilised FRP bridge deck<br />

– local buckling of the deck under<br />

stress vertical to the pultrusion<br />

direction.<br />

Tests for Technical Approval<br />

Overview<br />

The FRP deck ASSET used for the<br />

Frieberg Bridge was developed within<br />

the scope of an EU research project<br />

and had already been utilised for a<br />

smaller bridge in 2002. 9 From that project,<br />

material properties for the uniaxial<br />

load-bearing behaviour in the pultrusion<br />

direction were already available<br />

and applicable. Several additional specific<br />

values such as the shear strength<br />

of the FRP members and, moreover,<br />

the ultimate strength of the glue joint<br />

in tension and shear, the characteristics<br />

of the polymer concrete, the composite<br />

action between steel and FRP<br />

and the load-bearing behaviour due<br />

to concentrated wheel loads had to be<br />

determined.<br />

Composite Action between Steel<br />

and FRP<br />

Two main values are decisive in ensuring<br />

composite action: the strength of<br />

the adhesive joints between the steel<br />

girders and FRP deck panel, and the<br />

strength of the deck panel in compression<br />

vertical to the pultrusion direction.<br />

For the adhesive joints, tension<br />

and shear tests were preformed on<br />

small specimens. Rupture occurred<br />

within the FRP elements, and not<br />

within the adhesive or the interlayer.<br />

All tested specimens showed delamination<br />

within the top flange of the<br />

FRP deck. Therefore, it was of no<br />

importance to the ultimate strength,<br />

whether the surfaces were ground and<br />

degreased or not. The determined ultimate<br />

strength was far higher than the<br />

requirements identified in the analysis.<br />

The load-bearing behaviour of the<br />

deck under compression lateral to the<br />

Flange<br />

Inner web<br />

Outer web


pultrusion direction has already been<br />

described within the context of other<br />

research work. 6 Analogue experiments<br />

were conducted in order to confirm the<br />

values and secure a wider data basis.<br />

Therefore a 750 × 600 mm section of<br />

the deck was tested in an upright position<br />

under central load up to the point<br />

of rupture (Fig. 5). Because of local<br />

buckling, the strength was considerably<br />

lower than what could have been<br />

expected from the laminate architecture.<br />

The buckling appeared within the<br />

reinforcing layers of the overlap joint<br />

700<br />

between the individual elements. The<br />

failure mode became decisive for the<br />

design of the bridge. An optimization<br />

of the cross section, especially around<br />

the joining area could increase the<br />

compression strength perpendicular to<br />

the direction of pultrusion.<br />

Concentrated Wheel Loads<br />

Stress concentrations due to concentrated<br />

wheel loads are critical for the<br />

thin walls of the FRP deck. The polymer<br />

concrete used for the surfacing<br />

can bear tension and shear stresses.<br />

The surfacing system and the top<br />

flange of the FRP deck are assumed<br />

to be in composite action. Thus, the<br />

stress is distributed over a larger area.<br />

In addition, the polymer concrete surfacing<br />

reinforces the FRP deck flange<br />

(Fig. 6). This approach is similar to the<br />

design concept for orthotropic steel<br />

superstructures. The system is stable<br />

for wheel loads according to ENV<br />

1991–3.<br />

Construction and Assembly<br />

of the Superstructure<br />

Pre-Assembly in Shop<br />

Normal force on a 600 mm wide specimen (kN)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

The entire superstructure was preassembled<br />

in a storehouse located 20<br />

km from the building site. This allowed<br />

for protection against weathering and<br />

guaranteed optimal working conditions.<br />

The steel beams had already<br />

been welded when delivered and furnished<br />

with corrosion protection in the<br />

hall. The FRP deck had already been<br />

partially glued by the manufacturer<br />

and delivered in 5,00 × 1,50 m courses.<br />

The bonding of the deck panels onto<br />

the steel beams could be performed<br />

within 1 week. The FRP was ground<br />

and cleaned directly before the bonding;<br />

the adhesively bonded joints of<br />

the steel beams had already been<br />

sandblasted.<br />

0<br />

0 1 2 3 4 5 6 7<br />

Deformation (mm)<br />

Fig. 5: Test set-up and results for determination of compressive strength of FRP—bridge<br />

deck perpendicular to the direction of pultrusion<br />

Load on a area of 400 × 200 mm (kN)<br />

–175<br />

–150<br />

–125<br />

–100<br />

–75<br />

–50<br />

–25<br />

0<br />

0<br />

–1<br />

–2<br />

–3 –4 –5<br />

Deflection (mm)<br />

With surfacing<br />

Without surfacing<br />

Fig. 6: Test set-up and results for concentrated wheel loads, with and without surfacing<br />

layer<br />

–6<br />

–7<br />

–8<br />

On completion of the deck, the edge<br />

casings were affixed with a second<br />

layer of FRP sections and the front<br />

face was closed. After that a 45 mm<br />

thick layer of polymer concrete was<br />

applied onto the FRP. In the wet condition,<br />

the chosen coating was very<br />

stiff and debris-like and had to be<br />

levelled with appropriate tools to a<br />

nominal thickness. The treatment of<br />

the surfacing in one working sequence<br />

without joints proved to give the best<br />

results. The final jobs were the mounting<br />

of the railings, the edge sections at<br />

the transition between superstructure<br />

and abutment and the casting of 5 mm<br />

thick polymer surfacing on the edge<br />

casings.<br />

Transportation, Lifting and<br />

Positioning of the Superstructure<br />

Transporting the bridge to the building<br />

site had to be done in one night,<br />

as the public road next to the assembly<br />

hall could only be blocked for that<br />

long. The bridge was transported to the<br />

building site on a low platform trailer<br />

at a maximum speed of 50 km/h. The<br />

total weight of the superstructure was<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 403


Fig. 7: Mounting of superstructure on construction<br />

site (photo credit: Fiberline Composites<br />

A/S, Denmark)<br />

approximately 60 tons. The total width<br />

of the carriage was only 5 m; support<br />

vehicles were not necessary and the<br />

motorway was not closed. The superstructure<br />

was lifted the next morning<br />

within 2 h onto the abutments (Fig. 7)<br />

by two truck-mounted cranes, after<br />

which the prepared supporting pockets<br />

were grouted (Fig. 8). With this<br />

completed, only the finishing jobs on<br />

the abutments were pending before<br />

the bridge could be made operational.<br />

Conclusion<br />

The use of fibre composites has great<br />

potential in bridge building. The possibility<br />

of pre-assembling in the hall<br />

allows for rapid installation on site. In<br />

this project, the critical working stages<br />

of gluing the FRP panels, as well as<br />

the lifting and positioning of the<br />

superstructure were performed without<br />

any problems. The handling of the<br />

prefabricated products is relatively<br />

non-sensitive. Detailed recording<br />

and tracking the quality of important<br />

components and working stages in<br />

the tender documents proved to be<br />

worthwhile.<br />

Monitoring the bridge over the next<br />

few years should contribute to gathering<br />

experience and optimising building<br />

methods for FRP bridge decks. Paying<br />

Fig. 8: Superstructure being lowered shortly before settling down in the grouting—pocket<br />

(photo credit: ASV Gelnhausen, Germany)<br />

close attention to detailed specifications<br />

of technical details in the tender documents<br />

and during construction proved<br />

to be beneficial. Thus, the necessary<br />

execution excellence could be obtained<br />

and risks for the principal could be<br />

minimised. With good collaboration<br />

between the consultant and the public<br />

authority, and a high commitment from<br />

all parties involved, this project could<br />

be completed successfully.<br />

References<br />

[1] Kuhlmann U, Pelke E, et al. Ganzheitliche<br />

Wirtschaftlichkeitsbetrachtungen bei Verbundbrücken.<br />

Stahlbau 76 (Heft 2), Seite 105ff,<br />

Germany.<br />

[2] Webpage of Federal Highway Administration<br />

(USA). www.fhwa.dot.gov/Bridge/FRP (as on<br />

Aug.16, <strong>2010</strong>).<br />

[3] Lee SW, Kee-Jeung H. Experiencing More<br />

Composite-Deck Bridges and Developing<br />

Innovative Profile of Snap-Fit Connection.<br />

COBRAE Conference, Stuttgart, 2007.<br />

[4] Keller T, et al. Long-term performance of<br />

a glass fiber-reinforced polymer truss bridge.<br />

J. Compos. Constr. (ASCE) 01/2007, Vol. 11,<br />

pp.99–108.pp.99.<br />

[5] ENV 1991–3 Traffic Loads on Bridges.<br />

[6] Gürtler H. Composite Action of FRP Decks<br />

adhesively bonded to Steel Main Girders. PhD<br />

thesis No. 3135, Prof. Thomas Keller, EPFL<br />

Lausanne (CH), CCLab, 2004.<br />

[7] BÜV-Empfehlung. Tragende Kunststoffbauteile<br />

im Bauwesen (TKB). Bau-<br />

Überwachungsverein e.V., 2002 (Entwurf).<br />

[8] Clarke J, et al. Structural Design of Polymer<br />

Composites—EUROCOMP Design Code and<br />

Handbook. E&FN SPON: London, 1996.<br />

[9] Luke S, et al. Advanced composite bridge<br />

decking system—project ASSET. Struct. Eng.<br />

Int. 2002; 12: 76–79.<br />

<strong>SEI</strong> Data Block<br />

Owner:<br />

State of Hessen, Germany in the<br />

representation of Federal Republic<br />

of Germany<br />

Designer:<br />

Knippers Helbig Advanced<br />

Engineering, Stuttgart – New York<br />

Material tests and monitoring:<br />

Universität Stuttgart, Institute for<br />

Building Structures and Structural<br />

design (itke), Stuttgart (D)<br />

Design Construction Phase:<br />

KHP Planungsgesellschaft<br />

Main contractor:<br />

LS Bau GmbH & Co. KG, Giessen (D)<br />

FRP fabricator:<br />

Fiberline Composites A/S,<br />

Middelfart (DK)<br />

FRP (t): 13,5<br />

Span lengths (m): 27,0<br />

Construction cost (EUR million): 0,55<br />

Service date: August 2008<br />

404 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Current and Future Applications of Glass-Fibre-Reinforced<br />

Polymer Decks in Korea<br />

Sung Woo Lee, President, Prof., Kookmin University, Seoul, Korea; Kee Jeung Hong, Prof., School of Civil Eng. Kookmin<br />

University, Seoul, Korea; Sinzeon Park, Associate Manager, Kookmin Composite Infrastructure Inc., Kyungki-do, Korea.<br />

Contact: kjhong@kookmin.ac.kr<br />

Abstract<br />

In comparison to concrete or steel<br />

bridge decks, glass-fibre-reinforced<br />

polymer (GFRP) compos ite bridge<br />

deck is highly economical, since its<br />

lightweight property reduces initial<br />

construction cost for the foundation,<br />

and the high durability decreases<br />

the life-cycle cost for the bridges.<br />

Furthermore, the duration of construction<br />

is reduced significantly because of<br />

the short installation time of the lightweight<br />

GFRP composite decks.<br />

Korea is one of the leading countries<br />

in the construction of composite-deck<br />

bridges in recent years. In Korea, 13<br />

road bridges (deck area = 15 917 m 2 )<br />

have been constructed and six more<br />

road bridges (deck area = 9747 m 2 )<br />

will be constructed in the near future.<br />

Furthermore, 19 footbridges (deck area<br />

= 14 921 m 2 ) have been constructed to<br />

date and six more footbridges (deck<br />

area = 14 204 m 2 ) will be constructed in<br />

the near future. Among them, there are<br />

two remarkable projects: the Noolcha<br />

Bridge in Busan Newport which was<br />

constructed with the world’s largest<br />

composite-deck panel, which is 300 m<br />

long and 35 m wide, and the existing<br />

1,7 km-long walkway of the Hangang<br />

Bridge in Seoul which was expanded<br />

from a width of 2,5 to 5 m by replacing<br />

the existing concrete decks with composite<br />

decks.<br />

Keywords: GFRP; composite deck;<br />

road bridge; footbridge; walkway<br />

expansion; snap-fit.<br />

Introduction<br />

To cope with the problems of deterioration<br />

and corrosion in conventional steel<br />

and concrete materials, highly durable<br />

and lightweight fibre-reinforced polymer<br />

(FRP) composites are considered<br />

to be one of the most promising alternative<br />

materials for civil infrastructures.<br />

Among many applications of<br />

these materials, glass-fibre-reinforced<br />

polymer (GFRP) composite decks for<br />

bridges are particularly notable, since<br />

they have been used successfully to<br />

replace conventional concrete and<br />

steel decks on a number of bridges. The<br />

GFRP composite decks for bridges<br />

have significant advantages compared<br />

to conventional concrete and steel<br />

decks as they are highly durable and<br />

corrosion free. Much longer service<br />

lives and lower maintenance costs are<br />

expected for GFRP composite decks<br />

than for conventional concrete and<br />

steel decks, which will, in most cases,<br />

result in much lower life-cycle cost<br />

(LCC). The lightweight of GFRP composite<br />

decks reduces the dead load by<br />

as much as 80% compared to conventional<br />

concrete decks. Much slimmer<br />

substructures are possible for bridges<br />

on account of the use of lightweight<br />

GFRP composite decks. When a GFRP<br />

composite deck is used for re-decking<br />

of a bridge, the capacity for live load on<br />

the bridge is consequently upgraded<br />

without strengthening its substructures.<br />

Furthermore, since GFRP composite<br />

decks can be installed easily and<br />

quickly, the duration of construction<br />

reduces significantly and the amount of<br />

disruption to traffic lessens. This allows<br />

considerable savings in direct and indirect<br />

costs to the urban community.<br />

In view of these notable advantages<br />

of GFRP composite decks, numerous<br />

studies have been carried out and an<br />

increasing number of field applications<br />

have been reported. 1,2 Moreover,<br />

many profiles of GFRP composite<br />

decks have been developed and put<br />

into practice since the 1990s. 2 In recent<br />

times, Korea has become one of the<br />

leading countries in the construction<br />

of composite-deck bridges. In Korea,<br />

more than 30 bridges (road and footbridges)<br />

with GFRP composite decks<br />

have been constructed. The “Noolcha<br />

Bridge” in the Busan Newport area<br />

was constructed with the world’s largest<br />

(300 m long and 35 m wide) composite-deck<br />

panel, and the 1,7 km-long<br />

walkway of the “Hangang Bridge” in<br />

Seoul was expanded from a width of<br />

2,5 to 5 m by replacing the existing<br />

concrete decks with the composite<br />

decks. Many more bridges with GFRP<br />

composite decks are scheduled to be<br />

constructed.<br />

The GFRP composite decks with<br />

tongue-and-groove connections were<br />

developed and commercialised by the<br />

authors’ university and a private company.<br />

3 Subsequently, the snap-fit connections<br />

of GFRP composite decks<br />

have been invented to significantly<br />

improve the construction quality and<br />

reduce the installation time compared<br />

to tongue-and-groove connections of<br />

GFRP composite decks. 4–11<br />

Developed GFRP Composite<br />

Decks<br />

Tongue-and-Groove Composite<br />

Decks<br />

Through extensive studies such as flexural<br />

tests, compressive fatigue tests,<br />

flexural fatigue tests, shear connection<br />

tests, traffic barrier tests, pavement<br />

bonding test, accelerated chemical<br />

tests and field load tests, a compositedeck<br />

profile that uses a tongue-andgroove<br />

connection was developed. As<br />

shown in Fig. 1a, it has three trapezoidal<br />

cells 200 mm in height (TG200),<br />

and is fabricated by pultrusion. To<br />

fabricate the laminate of the deck,<br />

8800 Tex E-glass rovings are used in<br />

the longitudinal direction in conjunction<br />

with multi-axial stitched fabrics<br />

(90°, ±45°), and unsaturated polyester<br />

is used as a resin base. The deck is<br />

designed for typical girders spaced<br />

at 2,5 to 3,0 m under the DB24 truck<br />

load (front and rear wheel loads are<br />

24 and 96 kN, respectively) according<br />

to the Korean Highway Bridge Design<br />

Specifications. 12 The pultruded deck<br />

tubes are horizontally assembled by<br />

epoxy bonding to build deck panels for<br />

bridges as shown in Fig. 2a.<br />

Snap-Fit Composite Decks<br />

Use of the tongue-and-groove connection<br />

method is prevalent in the assembly<br />

of composite decks, but it causes<br />

two problems when used for bridge<br />

decks. For conventional construction<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 405


composite decks are rather limited<br />

and predictable. If the snap-fit composite<br />

decks are assembled without<br />

adhesive bonding, they can be easily<br />

disassembled and reused. The snap-fit<br />

connections also help to reduce the<br />

installation time and associated working<br />

costs.<br />

(a) Tongue-and-groove connection (TG200)<br />

Fig. 1: Pultrusion of composite decks<br />

Fig. 2: Connections for composite decks<br />

with concrete decks, shear connectors<br />

are installed on top of the girder prior<br />

to the placement of the concrete decks.<br />

For construction with the tongue-andgroove<br />

composite decks, in contrast,<br />

shear connectors cannot be installed<br />

before the placement and assembly of<br />

the composite decks on girders, since<br />

the composite decks are assembled by<br />

horizontal sliding on the top surface of<br />

the girder. Small holes at the locations<br />

of the shear studs must therefore be<br />

made on the composite decks prior to<br />

placing the composite decks on girders.<br />

After the placement and assembly<br />

of the composite decks, the shear studs<br />

are welded to the steel girders (or steel<br />

plates attached on the concrete girders)<br />

through small holes. Welding shear studs<br />

to girders through small holes is difficult<br />

and time-consuming and can result<br />

in poor welding quality. Accumulated<br />

horizontal gaps between the assembled<br />

composite decks become larger for a<br />

longer bridge. If the accumulated gaps<br />

are significantly large, it can cause mismatches<br />

between the locations of shear<br />

studs and the locations of the associated<br />

holes. This is the second problem that<br />

may produce construction errors and<br />

an increase in the construction time.<br />

To resolve the aforementioned<br />

problems related to conventional<br />

(a) Tongue-and-groove connection (TG200)<br />

(b) Snap-fit connection (SF200)<br />

(b) Snap-fit connection (SF200)<br />

tongue-and-groove connections, new<br />

profiles of composite decks with snapfit<br />

connections have been developed.<br />

The developed composite decks are<br />

assembled by mechanical snap-fitting<br />

in the vertical direction (Fig. 2b).<br />

Figure 1b shows the pultruded composite<br />

decks (SF200) with snap-fit<br />

connections for road bridges, which<br />

are not used yet in the field and are<br />

still under verification. Other snap-fit<br />

profiles have also been developed such<br />

as SF75L, SF75H, SF100 and SF125,<br />

which are not discussed in this paper<br />

explicitly. However, they are successfully<br />

applied in 19 footbridges in Korea<br />

in view of the easy assembly of the<br />

composite decks on the girder. The differences<br />

between these profiles lie in<br />

the height of the profile and the thickness<br />

of the flanges and webs.<br />

The newly developed snap-fit composite<br />

decks significantly improve<br />

workability and quality in welding<br />

shear studs to girders since the shear<br />

studs are welded with enough working<br />

space before placement of the composite<br />

decks. Furthermore, the developed<br />

decks help to avoid mismatches<br />

between the locations of shear studs<br />

and the locations of associated holes<br />

on the composite decks, since the<br />

horizontal gaps between the snap-fit<br />

Projects Examples<br />

Road Bridges<br />

The developed composite decks with<br />

tongue-and-groove connections at a<br />

height of 200 mm (Fig. 1a) have been<br />

applied for the construction of road<br />

bridges in Korea, with different girder<br />

types such as steel-plate girders, steelbox<br />

girders, pre-stressed concrete girders<br />

and reinforced concrete girders. 13<br />

road bridges (deck area = 15 917 m 2 )<br />

have been constructed and six more<br />

road bridges (deck area = 9747 m 2 )<br />

are planned to be constructed. Among<br />

these applications, the Bongsan Third<br />

Bridge and the Noolcha Bridge are<br />

shown in Figs. 3 and 4, respectively.<br />

The 36 m long and 7 m wide composite-deck<br />

panel was assembled on the<br />

steel-plate girders for the Bongsan<br />

Third Bridge as shown in Fig. 3a, construction<br />

on the bridge was completed<br />

in 2007. The Noolcha Bridge constructed<br />

in 2006 at the Busan Newport<br />

is considered a milestone since it has<br />

the world’s largest composite deck.<br />

Figure 4a shows the placing of the<br />

composite-deck panel from the storage<br />

site onto the RC girders using a crane<br />

for the Noolcha Bridge. Placing a single<br />

composite-deck panel (2 × 17,5 m)<br />

takes approximately 2 minutes, which<br />

is a significant reduction in construction<br />

time over that required for the<br />

placing of conventional concrete- or<br />

steel-deck panels. This quick installation<br />

of decks is a significant advantage<br />

in using the composite decks for<br />

bridge construction. Furthermore, the<br />

construction costs for the marine piles<br />

were significantly reduced because of<br />

the lightweight of the composite decks.<br />

Asphalt was used for the pavement of<br />

the Noolcha Bridge.<br />

Footbridges and Walkway Expansion<br />

The recently developed composite<br />

decks TG200, SF75L, SF75H, SF100<br />

and SF125 have been applied extensively<br />

for the construction of footbridges<br />

in Korea. These footbridges<br />

are constructed in short time periods<br />

as compared with conventional concrete-<br />

or wood-deck footbridges, since<br />

406 Technical Report Structural Engineering International 4/<strong>2010</strong>


(a) Placing GFRP composite decks (TG200)<br />

Fig. 3: Bongsan Third Bridge (steel-plate girder bridge)<br />

(b) Completed bridge<br />

of the Hangang Bridge in Seoul was<br />

upgraded with ease by replacing the<br />

existing concrete decks with the lightweight<br />

composite decks (SF75L). The<br />

existing walkway, which was 2,5 m in<br />

width, was expanded to 5 m through<br />

this upgrade without strengthening<br />

the substructures. After the existing<br />

concrete decks had been removed,<br />

Fig. 6a shows placing of the composite<br />

decks for the expanded walkway on<br />

both sides of the Hangang Bridge. The<br />

red lane for bikes and the grey lane<br />

for pedestrians are shown in Fig. 6b.<br />

Based on these successful applications<br />

of composite decks, more projects for<br />

new bridges and walkway expansions<br />

have been scheduled.<br />

Conclusion<br />

(a) Placing GFRP composite decks (TG200)<br />

Fig. 4: Noolcha Bridge (RC girder bridge)<br />

(a) Placing GFRP composite decks (SF100)<br />

Fig. 5: Wolchul Mountain Bridge (suspension bridge)<br />

(a) Placing GFRP composite decks (SF75L)<br />

the developed composite decks are<br />

much easier to install. 19 footbridges<br />

(deck area = 14 921 m 2 ) have been<br />

constructed and six more footbridges<br />

(deck area = 14 204 m 2 ) are planned<br />

for construction.<br />

(b) Completed bridge<br />

(b) Completed bridge<br />

(b) Completed walkway (four side-lanes)<br />

Fig. 6: Walkway expansion of Hangang Bridge (arch bridge and steel-plate girder bridge)<br />

In 2006, the developed composite<br />

decks (SF100) were quickly assembled<br />

at high elevation for a suspension footbridge,<br />

called the Wolchul Mountain<br />

Bridge. Shown in Fig. 5, it is 53 m long<br />

and 1 m wide. The 1,7 km-long walkway<br />

GFRP composite decks have been<br />

widely applied for many road bridges<br />

and footbridges in Korea. Among<br />

these GFRP deck bridges, there are<br />

two remarkable projects: the Noolcha<br />

Bridge in Busan Newport has been<br />

constructed with the world’s largest<br />

composite-deck panel and the<br />

Hangang Bridge in Seoul, has been<br />

expanded with ease from a width of 2,5<br />

to 5 m by replacing the existing concrete<br />

decks with the composite decks.<br />

The newly developed snap-fit connection<br />

of the composite decks resolves two<br />

problems encountered during construction<br />

with the tongue-and-groove composite<br />

decks: poor welding conditions<br />

for the connection of the shear studs and<br />

the possible mismatch between the locations<br />

of shear studs and the associated<br />

locations of pre-drilled holes on the<br />

composite decks. By snap-fitting composite<br />

decks without adhesive bonding,<br />

simplified assembly and disassembly of<br />

the composite decks are ensured. This<br />

enables their application in temporary<br />

bridges or temporary roads. It is anticipated<br />

that the newly developed snapfit<br />

connection will pave the way for far<br />

wider applications of GFRP composite<br />

decks in future, and that GFRP composite<br />

decks will be promising alternatives<br />

to c onventional concrete, steel or<br />

wood decks for bridges.<br />

References<br />

[1] DARPA. Advanced Composites for Bridge<br />

Infrastructure Renewal-Phase II Tasks 16-<br />

Modular Composite Bridge. Defense Advanced<br />

Research Projects Agency. Technical Report Vol.<br />

IV. USA, 2000.<br />

[2] Keller T. Use of Fiber Reinforced Polymers<br />

in Bridge Construction. Structural Engineering<br />

Documents 7. IABSE (International Association<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 407


for Bridge and Structural Engineering).<br />

Switzerland, 2003.<br />

[3] Lee SW. Development of High Durable, Light<br />

Weight and Fast Installable Composite Deck.<br />

MOCT R&D Report. Ministry of Construction<br />

and Transportation, Korea, 2004.<br />

[4] Lee SW. Fiber Reinforced Polymer Composite<br />

Bridge Deck of Tubular Profile Having Vertical<br />

Snap-Fit Connection. US Patent No. US 7, 131,<br />

161 B2, USA; 2006.<br />

[5] Lee SW, Hong KJ. Development of Composite<br />

Deck Connection for Pedestrian Bridge Using<br />

Korean Traditional Wooden Joint Method.<br />

KOSEF Research Report. Korean Science and<br />

Engineering Foundation, Korea, 2007.<br />

[6] Lee SW, Hong KJ. Experiencing more composite-deck<br />

bridge and developing innovative<br />

profile of snap-fit connections. Proceedings of<br />

COBRAE Conference. Stuttgart, Germany, 2007.<br />

[7] Lee SW, Hong KJ. Constructing bridges<br />

with glass-fiber reinforced composite decks.<br />

Proceedings of 4th International Structural<br />

Engineering and Construction. RMIT University,<br />

Melbourne, Australia, 2007.<br />

[8] Lee SW, Hong KJ. Evolution of innovative<br />

snap-fit connection for pultruded ‘Delta Deck’.<br />

Proceedings of European Pultrusion Technology<br />

Association. Rome, Italy, 2008.<br />

[9] Lee SW, Hong KJ. Experiencing more<br />

GFRP composite bridge decks for vehicular<br />

and pedestrian bridges. Proceedings of IABSE.<br />

Chicago, USA, 2008.<br />

[10] Lee SW, Hong KJ, Ketel J. Composite ‘Delta<br />

Deck’ of innovative snap-fit connection for new<br />

and rehabilitated footbridges. Proceedings of<br />

Footbridge 2008. Porto, Portugal, 2008.<br />

[11] Lee SW, Hong KJ. Development of<br />

Light-Weight Composite Deck with Snap-Fit<br />

Connection for Rigmat and Bridge Deck. MOCT<br />

R&D Report. Ministry of Construction and<br />

Transportation. Korea, 2009.<br />

[12] Korea Road and Transportation Association.<br />

Korean Highway Bridge Design Specifications.<br />

Ministry of Construction and Transportation,<br />

Korea, 2005.<br />

SED 7 - Structural Engineering Document on FRP<br />

IABSE<br />

This Structural Engineering Document reviews the progress made<br />

worldwide in the use of fibre reinforced polymers as structural<br />

components in bridges until the end of the year 2000. It includes<br />

application and research recommendations with particular reference<br />

to Switzerland and is aimed at both students and practising engineers,<br />

working in the field of fibre reinforced polymers, bridge design,<br />

construction, repair and strengthening.<br />

Topics:<br />

Overview and Classification, Fibres and Matrices, Tensile Elements,<br />

Structural Components and Systems, FRP-Reinforced Concrete – Stateof-the-Art,<br />

Fibre Reinforced Polymers – State of the Art in Repair and<br />

Strengthening, Fibre Reinforced Polymers – State-of-the-Art in Hybrid<br />

New Structures, Fibre Reinforced Polymers – State-of-the-Art in All-<br />

Composite New Structures, Design, Codes and Guidelines, Application<br />

Recommendations, Research, Requirements and Recommendations.<br />

Price: CHF 40 for Members, CHF 70 for Non-Members<br />

www.iabse.org/publications/onlinshop<br />

408 Technical Report Structural Engineering International 4/<strong>2010</strong>


Field Issues Associated with the Use of Fiber-Reinforced<br />

Polymer Composite Bridge Decks and Superstructures<br />

in Harsh Environments<br />

Louis N. Triandafilou, P.E., M.ASCE, Senior Structural Eng., Federal Highway Administration Resource Center, Baltimore, MD,<br />

USA and Jerome S. O’Connor, P.E., F. ASCE, Manager, Bridge Engineering Program, Dept. of Civil, Structural, Environmental<br />

Engineering, University at Buffalo, The State University of New York, Buffalo, NY, USA. Contact: lou.triandafilou@dot.gov<br />

Abstract<br />

There has been ample time for the<br />

United States to evaluate the use of<br />

fiber-reinforced polymer (FRP) composites<br />

to serve as bridge decks and<br />

superstructures under real-world environmental<br />

and operating conditions.<br />

This provides a great opportunity to<br />

weigh in on the decision to move forward<br />

with these materials. By studying<br />

the successes and failures, materials<br />

and techniques used for the design and<br />

construction of these bridges can be<br />

improved so that we can move closer to<br />

the goals of maintenance-free bridges<br />

and service lives exceeding 100 years.<br />

Though only a few of these structures<br />

are instrumented for structural health<br />

monitoring, there are a sufficient number<br />

of lessons that can be gleaned from<br />

the experiences of various owners during<br />

the fabrication, installation, operation,<br />

and inspection of these structures.<br />

By addressing these issues head-on,<br />

advocates will be better equipped to<br />

have direct dialogue with those in the<br />

profession who remain skeptical about<br />

the hidden potential for a construction<br />

material that is strong, light, and corrosion<br />

resistant.<br />

Keywords: FRP; deck; composite;<br />

bridge deck; superstructure; bridge.<br />

Introduction<br />

In the United States, public traffic has<br />

been carried on fiber-reinforced polymer<br />

(FRP) bridges since the installation<br />

of the first superstructure in 1996.<br />

A summary of these installations is<br />

available in the literature. 1 Bridge owners<br />

and researchers can benefit from<br />

these installations by considering their<br />

ongoing use as a real-world laboratory.<br />

Although the vast majority of the<br />

bridges constructed or rehabilitated<br />

with composites are considered a success,<br />

there have been some undesirable<br />

performance issues, such as debonding<br />

of wearing surfaces. Investigating<br />

these issues and openly sharing the<br />

experiences and findings with others<br />

should be encouraged. By making this<br />

knowledge available to a wide range<br />

of owners, engineers, and researchers,<br />

cooperation can be brought about in<br />

the development of the modifications<br />

to design and construction practices<br />

that will help advance the state of<br />

the art and the use of composites in<br />

practice.<br />

The scope of the discussion herein is<br />

restricted to 117 identified structures<br />

that have utilized FRP for the bridge<br />

deck or the superstructure. Although<br />

a few of these structures are too short<br />

to meet the federal definition of a<br />

bridge (at least 6,1 m), they are being<br />

considered here, since the lessons are<br />

the same, regardless of span length.<br />

Also included are the decks that have<br />

been taken out of service because of<br />

operational problems, and FRP superstructures<br />

that use concrete, timber, or<br />

asphalt for the deck or wearing surface.<br />

Experience from these bridges<br />

has provided lessons pertaining to the<br />

fabrication of the FRP, application of<br />

the wearing surface, and their longterm<br />

performance under harsh environmental<br />

conditions. Excluded from<br />

the discussion are pedestrian bridges,<br />

timber glue-laminated bridges that use<br />

FRP as a laminate, and FRP stay-inplace<br />

(SIP) forms for a concrete deck.<br />

Performance Summary<br />

Of the installations considered in this<br />

study, 95% are still in service and being<br />

monitored visually as part of regularly<br />

scheduled bridge inspections. Six have<br />

been taken out of service because of<br />

structural defects or failures of the<br />

deck or superstructure panels. An<br />

undisclosed number have also exhibited<br />

problems such as cracking and<br />

spalling of an adhesively bonded wearing<br />

surface.<br />

Structural Integrity<br />

It is often difficult to state conclusively<br />

what caused a particular structural<br />

failure. The six products taken out of<br />

service were from various manufacturers<br />

and manufacturing techniques.<br />

– Two types of deck were removed<br />

from the Salem Avenue bridge in<br />

Ohio (one vacuum-assisted resin<br />

transfer molded (VARTM), and<br />

one hand-laid sandwich section),<br />

because the FRP flange had separated<br />

from the FRP web core. There<br />

may have been contributing factors<br />

that led to failure of the decks after<br />

the bridge was placed in service, but<br />

an initiator of the problems seems to<br />

have been a manufacturing defect.<br />

The connection between the decks<br />

and the steel beams may have also<br />

played a role by allowing too much<br />

movement under live loads.<br />

– A detour bridge in Iowa was taken<br />

out of service and is not being used<br />

because it is not considered structurally<br />

sound. It is a VARTM sandwich<br />

section that separated between the<br />

FRP flange and web (top face skin<br />

and core). This could have been<br />

caused by a manufacturing defect<br />

or possibly by pounding from truck<br />

traffic along the leading edge after<br />

it was placed in service. The impact<br />

was apparently due to a slight height<br />

difference between the deck surface<br />

and the approach pavement.<br />

– In West Virginia, two corrugated<br />

core decks from the same manufacturer<br />

have been taken out of service<br />

because of structural integrity issues<br />

with the hand lay-up sandwich section.<br />

On one, clips between the deck<br />

and steel floor beam failed, allowing<br />

excessive movement that apparently<br />

led to a crushing of the web core of<br />

the deck. On the other, there were<br />

leaking joints between panels, and<br />

apparent debonding of the top face<br />

skin and crushing of the FRP core<br />

that resulted in “potholes”.<br />

– A pultruded FRP deck was removed<br />

from a bridge in Oregon owing to<br />

unsatisfactory performance of the<br />

support and connection details. The<br />

connection, which was designed to<br />

be rigid between the FRP deck and<br />

the steel structure, did not serve as<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 409


intended and led to a failure of the<br />

grout haunch and shear stud pockets,<br />

excessive flexure of the deck,<br />

and cracking of the wearing surface.<br />

In addition, there are two bridges (in<br />

California and New York) that are<br />

being monitored closely. These were<br />

constructed by different manufacturers<br />

but both have been repaired after<br />

debonding of the sandwich sections was<br />

discovered. Figure 1 shows a void under<br />

an edge of the top face skin of one<br />

bridge superstructure and Fig. 2 shows<br />

a core of the top face skin of the other.<br />

Thermal Behavior<br />

Although composites used in other<br />

industries have been thoroughly investigated<br />

and used in service at extreme<br />

temperatures, there has been limited<br />

research done on bridges at extreme<br />

temperatures. 2 Thus, the observation<br />

of in-service bridges has been useful<br />

in demonstrating performance under<br />

these harsh conditions.<br />

FRP behaves slightly differently from<br />

traditional materials at cold temperatures<br />

and during fluctuations in temperature.<br />

For instance, not only is the<br />

coefficient of thermal expansion different<br />

from steel or concrete, within the<br />

same FRP unit, it may be different longitudinally<br />

as compared to transversely.<br />

In addition to the extent of the rmal<br />

expansion/contraction being different,<br />

the rate at which the changes occur is<br />

also different. In general, this has not<br />

created a problem. The exception is<br />

that the interface between the FRP and<br />

the applied wearing surface has sometimes<br />

failed; thermal behavior may be<br />

involved, though conclusive studies have<br />

not been made. See below for further<br />

discussion on bonded wearing surfaces.<br />

Drivers in cold climates are aware that<br />

the surface of a bridge can become<br />

icy before the roadway does. This is<br />

because the surface temperature of the<br />

roadway is moderated by that of the<br />

earth below it. Bridges do not benefit<br />

from the heat that is stored in that thermal<br />

mass, and respond to drops in air<br />

temperature more quickly. FRP decks<br />

react the same way, but the changes are<br />

even more dramatic because they have<br />

little thermal mass when compared to<br />

concrete decks. While this is not known<br />

to have caused any problems to date, it<br />

is important to be aware that the rate<br />

of temperature change is not what drivers<br />

may be expecting. Figure 3 shows<br />

frost on the deck while the approaches<br />

remain clear and dry.<br />

top surface while the bottom surface<br />

remains relatively cool. This typically<br />

occurs during autumn or spring when<br />

there are cool nights, but daytime temperatures<br />

rise rapidly and the dark<br />

roadway warms up by absorbing solar<br />

radiation. The difference in temperature<br />

between the top and the bottom<br />

has been observed to be as much as<br />

25°C (77°F). Analysis has shown that<br />

thermally induced stresses can be higher<br />

than stresses caused by live loads from<br />

traffic. Several FRP structures exhibit<br />

this phenomenon, but none have suffered<br />

damage as a direct result.<br />

Wearing Surface<br />

The serviceability issue most often<br />

reported as a problem concerns the<br />

wearing surface. Although no problems<br />

have resulted from the use of a<br />

concrete or timber deck used with<br />

FRP beams, or with asphalt used as a<br />

wearing surface on FRP decks (except<br />

in a case where there was a structural<br />

integrity issue with the deck itself),<br />

there have been numerous reports of<br />

cracking and debonding of polymer<br />

concrete wearing surfaces that were<br />

applied to FRP decks. The figures<br />

below highlight some of the relevant<br />

concerns. In general, wearing surface<br />

problems can be categorized as follows:<br />

– material selection;<br />

– application;<br />

– structural integrity of the FRP<br />

substrate;<br />

– movement or flexibility of the FRP<br />

substrate.<br />

Material Selection<br />

Fig. 1: Top face skin of the sandwich<br />

superstructure debonded from the core<br />

(California)<br />

Fig. 2: A sample drilled from the top face<br />

skin of a sandwich panel exhibits dry<br />

fibers. This manufacturing flaw may have<br />

led to debonding of the face skin from the<br />

core (New York)<br />

Fig. 3: Frosting of the deck surface while the<br />

roadway remains relatively warm and dry<br />

The use of snow plows on bridges that<br />

have FRP decks has not been cited as<br />

a problem. As is typical with conventional<br />

decks, high spots in the deck<br />

surface may exhibit polishing and wear<br />

sooner than the rest of the deck.<br />

High temperatures observed in the<br />

field frequently result in the “hogging”<br />

of thick FRP superstructure panels. The<br />

resulting camber (arching) does not<br />

seem to create a problem as long as the<br />

wearing surface and its bond with the<br />

FRP can handle the shear and tensile<br />

stresses. Hogging is a thermal distortion<br />

that results from a temperature gradient<br />

through the depth of a section that<br />

results in an expansion of the warm<br />

Fig. 4: Cracking along a construction joint<br />

created between two FRP superstructure<br />

panels. Use of a flexible material such as<br />

two-part silicone may have been a better<br />

material choice<br />

Figure 4 shows a construction joint<br />

between FRP superstructure panels<br />

that was filled with adhesive during<br />

installation. The wearing surface over<br />

the joint has broken up, making the<br />

joint susceptible to moisture intake.<br />

410 Technical Report Structural Engineering International 4/<strong>2010</strong>


Figure 8 shows uneven wear of a<br />

thin epoxy concrete wearing surface.<br />

Although snow plows can be expected<br />

to polish high spots on any deck,<br />

these areas are most likely due to<br />

improper proportioning of aggregate<br />

and resin, or inadequate mixing during<br />

application.<br />

Fig. 5: Precautions must be taken when<br />

using asphalt for movable bridges because<br />

the weight of the material can cause a failure<br />

of the bond when the bridge is raised<br />

Figure 5 shows a movable bridge that<br />

is typically an attractive structure on<br />

which to use a lightweight deck. It<br />

serves as a reminder that the wearing<br />

surface sometimes needs to be able to<br />

withstand gravity loads in addition to<br />

being able to fulfill its other functions.<br />

When the bridge is in the up position,<br />

shear stresses between the deck and<br />

the wearing surface material need to<br />

be accommodated so that the material<br />

does not slide off the structure.<br />

On particularly hot days, the shear<br />

capacity of the bond and/or the material<br />

may become reduced. One agency<br />

has had this experience.<br />

Fig. 6: The debonding of this 7 mm wearing<br />

surface may have been due to improper<br />

material selection<br />

Figure 6 shows a partial debonding of<br />

a thin (6 mm) wearing surface. In this<br />

instance, a trial application was made<br />

using an acrylic-modified Portland<br />

cement based material. Most of the<br />

bonded surface survived but areas<br />

where it was thinly applied peeled off<br />

within a few years.<br />

Application<br />

Fig. 7: Debonding, possibly from<br />

inadequate surface preparation<br />

Figure 7 illustrates a thin polymer<br />

overlay that spalled due to poor surface<br />

preparation and installation.<br />

Although this illustration is that of a<br />

concrete deck, the result is typical of<br />

what can result on an FRP deck.<br />

Fig. 8: Polishing due to improper mixing<br />

and application. Durability may have also<br />

been reduced due to selection and use of an<br />

aggregate that was too small<br />

Fig. 9: Polymer concrete applied over a<br />

field joint between FRP panels. Insufficient<br />

mixing may be contributing to premature<br />

deterioration<br />

Figure 9 shows deterioration of a<br />

polymer concrete over a joint. Field<br />

applications such as this are difficult<br />

because of the additional variables<br />

involved (human and environmental<br />

factors).<br />

Fig. 10: An attempt to create an integral<br />

polymer concrete wearing surface during<br />

deck fabrication still resulted in surface<br />

degradation<br />

Figure 10 shows a superstructure that<br />

had the wearing surface made integrally<br />

with the FRP panel as part of<br />

the fabrication process. This was done<br />

by adding aggregate to the resin while<br />

it was applied by hand to the top of<br />

the panel. This integral surface was<br />

intended to circumvent the possibility<br />

of debonding of the wearing surface.<br />

Although this strategy was partially<br />

successful, it appears that the surface<br />

raveled because there was a partial<br />

cure in place before the top materials<br />

were applied.<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 411


Figure 13 shows an example of a failing<br />

FRP panel that has been patched<br />

over. In this case, an epoxy resin was<br />

used; however, asphalt has also been<br />

used on decks where the depression<br />

(or pothole) was deep enough.<br />

Fig. 11: Polymer concrete spalling due to<br />

deviance from installation procedures<br />

Figure 11 illustrates a typical debond<br />

that exposes the FRP surface to ultraviolet<br />

(UV) light and wheel loads. The<br />

edges are also susceptible to water intrusion<br />

that will cause the hole to propagate.<br />

Fig. 12: Debonding of brittle epoxy wearing<br />

surface in cold temperatures<br />

Figure 12 is a picture of a 10 mm epoxy<br />

wearing surface that debonded during<br />

its first winter in service. The bond failure<br />

was attributed to improper surface<br />

preparation during installation but, as<br />

can be seen, the material broke up in a<br />

brittle fashion when it was debonded.<br />

The material’s lack of flexibility at low<br />

temperatures probably caused this.<br />

Water intrusion and its subsequent<br />

expansion upon freezing also accelerated<br />

the failure of the surface.<br />

Structural integrity of the FRP<br />

substrate<br />

Fig. 14: Random cracking over a<br />

failing sandwich superstructure panel.<br />

(Fig. 2 is a close-up of the core that<br />

was taken)<br />

Figure 14 shows another type of surface<br />

failure. This depression resulted<br />

from a softening or crushing of the<br />

FRP web core where there was water<br />

intrusion. The wearing surface is not<br />

intended to bridge defects such as this<br />

and exhibits alligator cracking. Bridge<br />

inspectors need to be aware of this<br />

condition since maintenance workers<br />

are likely to treat it as a pothole and<br />

just patch over it.<br />

Movement or flexibility of the FRP<br />

substrate.<br />

Fig. 15: Cracking has occurred over<br />

bond lines of pultruded sections that<br />

were joined in the factory. Selection<br />

of the adhesive may have played<br />

a role<br />

Figure 15 illustrates reflective cracking<br />

that has occurred over factory bond<br />

lines between pultrusions.<br />

Fig. 16: Cracking over a transverse steel<br />

floor beam, probably from stresses caused<br />

by flexure of the FRP panel over the top<br />

flange. Water intrusion was also observed<br />

Figure 16 illustrates tensile cracking<br />

that can occur over the edges of a floor<br />

beam as the more flexible FRP deforms<br />

under wheel loads and thermal stressing.<br />

In this case, the deck is secured to<br />

the top flange of the floor beam below<br />

by tightly bolting them together.<br />

Examples of successful wearing<br />

surfaces<br />

Fig. 17: Although asphalt wearing surfaces<br />

are not lightweight, they have worked well<br />

in many instances. No special bonding<br />

layer was applied to the FRP<br />

Figure 17 shows an FRP superstructure<br />

that was installed and then overlaid with<br />

asphalt. This approach has been used<br />

repeatedly with success. This adds a substantial<br />

amount of weight and detracts<br />

one of the benefits of using FRP (its<br />

lightweight nature). The additional dead<br />

load will also need to be addressed with<br />

respect to its contribution toward creep<br />

and the potential for brittle rupture.<br />

Although asphalt may not be desirable<br />

in all situations, there are some situations<br />

where it may be the best alternative.<br />

It also has the benefit of being a<br />

well-known material that maintenance<br />

crews are comfortable with.<br />

Fig. 13: This deficiency (which has been<br />

patched) is due to a failing sandwich<br />

superstructure panel rather than the<br />

material or methods used for the wearing<br />

surface.<br />

412 Technical Report Structural Engineering International 4/<strong>2010</strong>


Fig. 18: Thin polymer concrete surfaces<br />

provide good skid resistance and protection<br />

with little extra weight<br />

Figure 18 illustrates a successful installation<br />

of a polymer concrete overlay.<br />

After 5 years of service, this deck surface<br />

is still in “like-new” condition.<br />

New York State and Oregon have<br />

undertaken research projects to<br />

explore some of the wearing surface<br />

issues presented above. 3–5 The New<br />

York report by Aboutaha recommends<br />

applying the surface as two courses, one<br />

specifically designed to insure a good<br />

bond and the upper course to provide<br />

the requisite durability and skid resistance.<br />

3 Reports of any such installations<br />

are yet to be made available.<br />

Conclusion<br />

The vast majority of FRP deck and<br />

superstructure installations have performed<br />

well structurally, but success<br />

in the field has been diminished by the<br />

poor performance of several wearing<br />

surfaces, especially the ones made of<br />

thinly applied bonded materials that<br />

were used for weight saving. Better<br />

performance in the future can be<br />

obtained by proper selection of material,<br />

meticulous adherence to wellplanned<br />

installation procedures, use of<br />

pull-off tests to check for proper surface<br />

preparation and bond, and use of<br />

a two-course system, where one is used<br />

to ensure bonding and a top layer is<br />

used for skid resistance and durability.<br />

References<br />

[1] Triandafilou L, O’Connor J. FRP Composites<br />

for Bridge Decks and Superstructures: State of<br />

the Practice in the U.S. Proceedings of International<br />

Conference on Fiber Reinforced Polymer<br />

(FRP) Composites for Infrastructure Applications,<br />

University of the Pacific, Stockton, CA,<br />

2009.<br />

[2] Dutta P, Kwon S. Fatigue Performance<br />

Evaluation of FRP Composite Bridge Deck<br />

Prototypes Under High and Low Temperatures.<br />

82nd Annual Meeting, Compendium of Papers<br />

on CD ROM, Transportation Research Board,<br />

Washington, D.C., 2003.<br />

[3] Aboutaha R. Investigation of Durability<br />

of Wearing Surfaces for FRP Bridge Decks,<br />

Syracuse University, 2008.<br />

[4] Wattanadechachan P, Aboutaha RS, Hag-<br />

Elsafi O, and Alampalli S. Thermal compatibility<br />

and durability of wearing surfaces on FRP bridge<br />

decks. ASCE Journal of Bridge Engineering<br />

2006; 11(4), pp. 465-473; July–August.<br />

[5] Barquist G, Lovejoy S, Nelson S, Soltesz S.<br />

Evaluation of Wearing Surface Materials for<br />

FRP Bridge Decks OR-DF-06-02, Oregon<br />

Department of Transportation, 2005.<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 413


Examples of Applications of Fibre Reinforced Plastic<br />

Materials in Infrastructure in Spain<br />

Anurag Bansal, Research Engineer; John F. Monsalve Cano, Research Engineer; Bladimir O. Osorio Muñoz, Research<br />

Engineer; Carlo Paulotto, Dr., Research Engineer; Acciona Infraestructuras, Alcobendas, Spain. Contact: carlo.paulotto@acciona.es<br />

Abstract<br />

The introduction of new materials<br />

such as fibre reinforced plastics (FRPs)<br />

into construction practices has been<br />

discussed in this paper. Hereafter,<br />

some of the most significant examples<br />

of the applications of these new materials<br />

are presented in some detail.<br />

Keywords: carbon; glass; fibre; matrix;<br />

reinforcement.<br />

Introduction<br />

For the sake of clarity, the examples<br />

concerning the application of FRP<br />

materials in civil infrastructure in this<br />

paper are divided into two groups:<br />

structures and structural members<br />

entirely made with FRP materials and<br />

concrete structures or structural members,<br />

where FRP materials are used as<br />

external or internal reinforcement. The<br />

FRP beams employed to build up the<br />

girders of a road overpass along the A8<br />

freeway close to the Asturias Airport<br />

in the north of Spain and those used<br />

for the girders of two bridges along the<br />

M-111 freeway in Madrid belong to<br />

the former group. To the same group<br />

belong two FRP slabs, used as a channel<br />

bed during the channelling work of<br />

the Rio Mesoiro River in Galicia and<br />

the spiral stairs designed and manufactured<br />

for placement in a pumping well<br />

along one of the metro lines in Madrid.<br />

The FRP sheets used to reinforce both<br />

in shear and in bending a “reshaped”<br />

concrete beam in a building in Aragón<br />

and the FRP reinforcing bars used in<br />

some parts of the concrete substructure<br />

of the new tramway in Granada belong<br />

to the latter group.<br />

Airport in the north of Spain [1]. This<br />

is a four-span bridge of total length<br />

of 46 m (see Fig. 1a). The bridge girders<br />

are three continuous longitudinal<br />

carbon fibre beams on three intermediate<br />

supports connected to the<br />

reinforced concrete deck through<br />

alkali-resistant glass fibre shear connectors<br />

(see Fig. 1b). The beams are<br />

characterized by trapezoidal cross<br />

sections. They were manufactured in<br />

two parts and transported to the work<br />

site where they were joined by means<br />

of adhesive step joints. Each step joint<br />

was manufactured by placing the two<br />

parts of a beam side-by-side, wrapping<br />

them with pre-impregnated carbon<br />

fabric which was finally consolidated<br />

using the vacuum bag system. Using a<br />

crane, it took only half a day to place<br />

the three 46 m long beams on their<br />

supports. This was possible thanks<br />

to the minimal weight of the beams<br />

(a)<br />

at 1 kN/m. Glass fibre stay-in-place<br />

formworks were used to cast the concrete<br />

deck. This bridge was the result<br />

of an international research project<br />

involving Spain, Colombia and Cuba,<br />

and it was aimed at demonstrating the<br />

possibility of building a road overpass<br />

using FRP materials.<br />

The other two bridges are located<br />

on the outskirts of Madrid along the<br />

M-111 freeway [2]. These two bridges<br />

are identical, each made up of three<br />

simply supported spans (10, 14 and<br />

10 m) with a 20,40 m wide box-girder<br />

deck (see Fig. 2a). The four hybrid<br />

carbon–glass fibre deck beams have<br />

reverse “Ω”-shaped cross sections<br />

(see Fig. 2b). Each beam was closed by<br />

connecting its top flange to a sandwich<br />

panel along its entire length. This panel<br />

initially acts both as the top flange of<br />

the beam and as formwork for the<br />

Fig. 1: (a) General view of the bridge in Asturias; (b) View of the carbon fibre beams and<br />

the glass fibre stay-in place formworks<br />

(b)<br />

Structures and Structural<br />

Members Made Entirely<br />

of FRP Materials<br />

Asturias and M-111 Bridges<br />

Since 2003, three vehicular bridges<br />

have been designed, manufactured<br />

and erected whose girder beams are<br />

completely made of FRP materials.<br />

The first bridge is located along<br />

the highway leading to the Asturias<br />

(a)<br />

(b)<br />

Fig. 2: (a) General view of one of the twin bridges along the M-111 freeway; (b) One of<br />

the FRP beams during positioning<br />

414 Technical Report Structural Engineering International 4/<strong>2010</strong>


uncured concrete deck slab. Once the<br />

concrete has strengthened, the slab<br />

contributes to the deck strength, as it<br />

is connected to the top flanges of the<br />

beams through steel studs. The minimal<br />

weight of the beams (3 kN/m)<br />

meant that they could be placed on the<br />

supports using a simple crane truck.<br />

12 beams were positioned during<br />

one working day. These two bridges<br />

were developed as demonstrators in<br />

the framework of a research project<br />

funded by the European Commission.<br />

The overall objective of the project<br />

was the development of a new highperformance<br />

and cost-effective construction<br />

concept for bridges based on<br />

the application of rapid-renewal and<br />

long-life service infrastructures in the<br />

countries that joined the European<br />

Union in 2004.<br />

lay-up and consolidated using the<br />

vacuum bag technique. The slabs were<br />

2,50 m wide and 14,50 m long, and had<br />

a thickness of 0,05 m (see Fig. 3). They<br />

were manufactured in a workshop in<br />

Madrid and moved to the work site by<br />

truck. After the two FRP bottom slabs<br />

were put in place, the lateral walls<br />

and top slabs of reinforced concrete<br />

were cast (see Fig. 4). The structural<br />

connection between the FRP bottom<br />

slabs and the lateral walls was ensured<br />

using steel studs with one of their ends<br />

embedded in the FRP slabs.<br />

Spiral Stairs for a Pumping Well<br />

in Madrid<br />

The stairs of one of the pumping<br />

wells along one of the metro lines<br />

of Madrid were exposed to a highly<br />

aggressive environment owing to<br />

water and gasoline infiltrations. The<br />

well has a depth of 21,15 m and a<br />

Rio Mesorio River Channelling<br />

During the channelling work of the<br />

Rio Mesorio River in the Spanish<br />

region of Galicia, the channel had<br />

to cross a pre-existing oil pipeline at<br />

one point [3]. According to the original<br />

design, the channel consisted of<br />

a reinforced concrete hollow section<br />

composed of two identical cells. The<br />

section was 2,40 m high and 5,40 m<br />

wide, with all the walls having the<br />

same thickness of 0,30 m, except for<br />

the central vertical diaphragm that<br />

had a thickness of 0,50 m. During the<br />

excavation phase, it was found that<br />

the thickness of the soil layer covering<br />

the pipeline was less than expected.<br />

The level of the top face of the channel’s<br />

bottom slabs however could not<br />

be varied, but executing the original<br />

plan would have meant interference<br />

between the bottom slabs and the oil<br />

pipes. To cope with this problem, the<br />

only feasible solution appeared to<br />

be reduction of the thickness of the<br />

slabs constituting the channel bed by<br />

replacing them with a material with<br />

better mechanical properties as compared<br />

with reinforced concrete. The<br />

use of steel was judged inappropriate<br />

due to possible corrosion problems.<br />

On the other hand, FRP materials,<br />

having both better mechanical properties<br />

than reinforced concrete and not<br />

suffering from corrosion phenomena,<br />

appeared to be the best alternative. To<br />

optimize the cost of the FRP slabs, it<br />

was decided to manufacture them with<br />

a mix of carbon and glass fibre fabrics<br />

that were impregnated with epoxy<br />

resin to overcome the constraint<br />

imposed by the carbon fibre sizing.<br />

The slabs were manufactured by wet<br />

Fig. 3: Positioning of the FRP channel bed slabs<br />

Fig. 4: View of the completed channel section<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 415


3,98 m side square cross section,<br />

whereas the access to the well from<br />

the ground level is 1,00 × 0,80 m only.<br />

Originally, these stairs which were<br />

placed in the well to reach the pumps<br />

for their routine maintenance were<br />

made of steel, but intensive corrosion<br />

soon made it dangerous for the<br />

workers and the decision to replace<br />

them was made in November 2008.<br />

Initially, the plan was to substitute<br />

the steel stairs with aluminium or<br />

stainless steel stairs, but the owner<br />

judged both solutions too expensive<br />

and expressed concerns regarding<br />

possible problems in the construction<br />

of the new stairs because of the<br />

limited room available around the<br />

well entrance. In fact, the entrance is<br />

located in front of one of the Madrid<br />

airport terminals between a ramp and<br />

a belt conveyor, partially covered by<br />

a shelter. A solution of spiral stairs<br />

with the mezzanine floors completely<br />

made of FRP materials was then proposed<br />

(see Fig. 5). This solution had<br />

the advantage that the FRP material<br />

used was lightweight and corrosion<br />

free, thanks to its inherent properties.<br />

Glass fibre was chosen as the<br />

reinforcing fibre in order to bring<br />

down the cost of the stairs as much<br />

as possible, and phenolic resin was<br />

selected as the matrix due to its fireretardant<br />

characteristic. Moreover, it<br />

was proposed that a modular solution<br />

with the different stair steps<br />

progressively mounted one on top of<br />

the other to form the stairs, be used.<br />

This allowed the different parts of<br />

the stairs to be moved into the well<br />

Fig. 5: View of the FRP spiral stairs for the<br />

metro pumping well<br />

from ground level by simply using a<br />

pulley. In fact, a single stair step and a<br />

mezzanine floor have weights of 0,30<br />

and 1,47 kN, respectively. The different<br />

parts of the stairs were manufactured<br />

using different techniques. The<br />

57 steps characterized by a sandwich<br />

structure with a fire- resistant core<br />

were manufactured through resin<br />

transfer moulding (RTM). The five<br />

landings have the same sandwich<br />

structure of the steps and were manufactured<br />

by resin infusion. The three<br />

mezzanine floors were manufactured<br />

by resin infusion, while their supports,<br />

which were finally connected by steel<br />

dowels to the concrete walls of the<br />

well, were manufactured by hand<br />

lay-up. The design, manufacturing<br />

and construction of the stairs took 2<br />

months.<br />

FRP Materials as<br />

Reinforcement for Concrete<br />

Structures<br />

Beam Reshaping<br />

Because of a mistake during the<br />

design phase, a 11m span beam<br />

belonging to the reinforced concrete<br />

supporting structure of a building<br />

in the Spanish region of Aragon<br />

had an incorrect depth of 1,37 m<br />

instead of 1,07 m. This beam was<br />

positioned along the border of the<br />

building and its extra depth affected<br />

the height of a window beneath the<br />

beam, running along its whole length.<br />

The beam was initially reinforced in<br />

tension with eight steel rebars having<br />

a diameter of 20 mm and placed<br />

in two layers near the bottom of the<br />

beam cross section. The shear reinforcement<br />

consisted of 10 mm diameter<br />

steel stirrups spaced at 0,30 m. It<br />

was initially thought the beam depth<br />

could be reduced by cutting off 0,30<br />

m of the bottom of the beam and bolting<br />

steel plates to the reshaped beam.<br />

The heavy weight of the steel plates,<br />

however, would have represented a<br />

serious problem during the strengthening<br />

phase, because of the necessity<br />

of lifting them to the beam level and<br />

keeping them in position during the<br />

bolting phase. All these operations<br />

would have been further complicated<br />

by the provisional supports that<br />

needed to be used to sustain the beam<br />

after removal of the tensile reinforcement.<br />

For this reason, the solution of<br />

reinforcing the reshaped beam with<br />

carbon fibre plates was considered<br />

a better option. After cutting the<br />

beam, the cut surface was cleaned by<br />

blowing it with compressed air and a<br />

primer was applied on it. The cut, dry<br />

carbon fabrics were applied to the<br />

bottom and sides of the beam and<br />

impregnated with epoxy resin. At the<br />

bottom of the beam, the carbon fibre<br />

filaments were mainly oriented along<br />

the beam’s axis to resist the tensile<br />

stresses induced by bending, while at<br />

the beam-sides the carbon fibre filaments<br />

were oriented at ±45° with<br />

respect to the beam’s axis to sustain<br />

the external shear forces.<br />

New Tramway in Granada<br />

During the construction of a new<br />

tramway in Granada in 2009, the<br />

owner requested that steel rebars<br />

in some portions of the reinforced<br />

concrete slab supporting the rails be<br />

removed. The reason for this request<br />

was that normal steel rebars interfered<br />

with the electromagnetic system<br />

that detected the position of<br />

the tramcars along the line. To solve<br />

this problem, it was decided to take<br />

advantage of the electromagnetic<br />

transparency exhibited by glass fibre<br />

reinforcing bars. Owing to the permanent<br />

character of this application,<br />

vinyl ester resin was employed<br />

as matrix for the rebars to protect<br />

the glass filaments from the alkaline<br />

environment in the Portland concrete.<br />

A total of 13 560 m of 16 mm straight<br />

rebars were manufactured by pultrusion<br />

and 26 448 m of 12 mm diameter<br />

stirrups were produced through a<br />

process similar to pultrusion. A large<br />

experimental test campaign was carried<br />

out to first evaluate the thermomechanical<br />

properties of the rebar to<br />

be used in the design phase and then<br />

to ensure that these properties were<br />

maintained over all the rebar lots that<br />

were delivered.<br />

Conclusion<br />

From the applications presented in<br />

this article, it is evident that it is possible<br />

to take advantage of the inherent<br />

properties of composite materials,<br />

such as corrosion resistance, minimal<br />

weight, high tensile strength and<br />

electromagnetic transparency, and to<br />

cope with scenarios in which the use<br />

of the traditional construction materials<br />

would be impossible or imply<br />

higher costs in terms of money or<br />

time.<br />

416 Technical Report Structural Engineering International 4/<strong>2010</strong>


References<br />

[1] Gutierrez E, Primi S, Mieres JM, Calvo I.<br />

Structural testing of a vehicular carbon fiber<br />

bridge: quasi-static, and short-term behavior. J.<br />

Bridge Eng. ASCE 2008; 13 (3): 271–281.<br />

[2] Primi S, Areiza M, Bansal A, Gonzalez A.<br />

New design and construction of a road bridge<br />

in composites materials in Spain: sustainability<br />

applied to civil works. In Proceedings of the 9th<br />

International Symposium On Fiber-Reinforced<br />

Polymer Reinforcement for Concrete Structures,<br />

2009. Oehlers DJ, Griffith MC, Seracino R<br />

(eds). Sydney, New South Wales, 2009, 62. ISBN:<br />

9780980675504.<br />

[3] Paulotto C, Primi S, Bansal A. A composite<br />

bed for the Rio Mesorio river. Proceedings of<br />

the Composites UK 10th Annual Conference,<br />

Birmingham, England, <strong>2010</strong>.<br />

<strong>SEI</strong> Data Block<br />

Owner:<br />

Ministerio de Fomento<br />

(for the Asturias bridge);<br />

Comunidad de Madrid<br />

(for the M111 bridges)<br />

Designer:<br />

Acciona Infraestructuras<br />

(for all the 3 bridges)<br />

Main contractor:<br />

Acciona Infraestructuras<br />

(for all the 3 bridges)<br />

FRP fabricator:<br />

Acciona Infraestructuras<br />

(for all the 3 bridges)<br />

FRP (t): 15 (for the Asturias bridge);<br />

22,5 (for each one the M111 bridges)<br />

Span lengths (m): 46 (for the Asturias<br />

bridge); 34 (for each one of the M111<br />

bridges)<br />

Construction cost (USD million): 0,5<br />

(for the Asturias bridge); 0,5 (for each<br />

one of the M111 bridges)<br />

Service date: 2004 (for the Asturias<br />

bridge); 2008 (for both the M111<br />

bridges)<br />

Sign up for an Institutional Subscription<br />

2 Print Copies per issue (Feb, May, Aug., Nov)<br />

Online Access to the Journal for the entire<br />

Institution via Site License<br />

Online Access to Current + Back issues (1991- )<br />

at no extra cost (1500 and more <strong>SEI</strong> articles)<br />

only CHF 630 a year.<br />

www.iabse.org/<strong>SEI</strong><br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 417


Fiber-Reinforced Polymer Decks for Movable Bridges<br />

Raymond D. Bottenberg, P.E., Oregon Department of Transportation, Bridge Preservation Engineering, Oregon, USA.<br />

Contact: raymond.d.bottenberg@odot.state.or.us<br />

Abstract<br />

Fiber-reinforced polymer (FRP)<br />

bridge decks are ideal for re-decking<br />

movable bridges. They can match the<br />

weight of existing steel grid decking<br />

to minimize bridge balance adjustment.<br />

They are bicycle- and motorcycle-friendly<br />

and quieter than steel<br />

grids. They protect the structure<br />

from precipitation and are corrosion<br />

resistant.<br />

In 2002, the Oregon Department of<br />

Transportation (ODOT) installed<br />

FRP decks on two movable bridges<br />

near Astoria. Cementitious grout<br />

pads supported the deck on the steel<br />

stringers, and the deck was secured<br />

to the stringers by welded steel studs<br />

in cementitious grout pockets. The<br />

wearing surface was to have asphalt<br />

concrete, but was changed to epoxy<br />

polymer concrete after the asphalt<br />

slid off a bascule leaf during a prolonged<br />

test lift. The grout details had<br />

failed on one of these bridges, resulting<br />

in deck replacement, and both<br />

bridges had experienced distress in<br />

the wearing surface. The attachment<br />

details and wearing surfaces used on<br />

these bridges did not accommodate<br />

the higher deflections associated with<br />

FRP decks, and were regarded as<br />

unacceptable.<br />

In 2005, the ODOT installed a FRP<br />

deck on a movable bridge in Florence.<br />

Neoprene sheets supported the deck<br />

on the steel stringers, and the deck<br />

was secured to the stringers with large<br />

structural blind fasteners installed<br />

from below. The wearing surface was<br />

urethane polymer concrete, placed<br />

by the broom and seed method. This<br />

installation showed none of the distress<br />

experienced on the two bridges at<br />

Astoria. The strong but flexible attachment<br />

details and wearing surface used<br />

on this bridge accommodate deflections<br />

well and were recommended for<br />

future installations.<br />

Keywords: bascule; movable bridge;<br />

pultruded FRP decking; steel<br />

grid decking; polymer concrete;<br />

structural blind fastener; staged<br />

construction.<br />

Introduction<br />

Historically, movable bridges have<br />

carried many kinds of decks including<br />

timber, reinforced concrete, welded<br />

steel grid, and orthotropic steel. One<br />

of the key properties of a movable<br />

bridge deck is its weight, which on<br />

both bascule-type bridges and vertical<br />

lift bridges directly affects the amount<br />

of counterweight needed and the span<br />

drive machinery power requirements.<br />

Often corrosion resistance is also<br />

important, as many movable bridges<br />

exist in corrosive marine environments.<br />

In 1997, the first pultruded fiber-reinforced<br />

polymer (FRP) bridge deck<br />

was produced by bonding tubular<br />

sections into panels. FRP decks have<br />

been installed on approximately 112<br />

bridges in the United States. In many<br />

cases, FRP bridge decking can closely<br />

match the weight and thickness of<br />

steel grid deck products. For movable<br />

bridges, this makes it an ideal deck retrofit<br />

material, as little adjustment of<br />

counterweights is required and little<br />

or no adjustment of roadway height<br />

is needed. Additional benefits include<br />

the inherent corrosion resistance of<br />

FRP, protection of the bridge structure<br />

from precipitation and road run-off,<br />

quiet ambient noise levels, and a more<br />

drivable surface for all vehicles, in particular<br />

motorcycles and bicycles.<br />

The ODOT rehabilitated two basculetype<br />

movable bridges in 2000 through<br />

2002, replacing deteriorated timber<br />

decks with FRP deck panels. Both<br />

bridges exist in a corrosive marine environment<br />

near Astoria, Oregon, and the<br />

FRP deck was selected to help prevent<br />

corrosion. Similarly in 2005, ODOT<br />

replaced a failing steel grid deck of<br />

a bascule-type movable bridge with<br />

an FRP deck. This bridge also exists<br />

in a corrosive marine environment in<br />

Florence, Oregon, and the FRP deck<br />

was selected to help prevent corrosion.<br />

ODOT Lewis and Clark<br />

River Bridge<br />

The Lewis and Clark River Bridge<br />

is a single-leaf bascule-type movable<br />

bridge built in 1924, carrying<br />

two traffic lanes on a 6,096 m wide<br />

roadway, which is supported by steel<br />

stringers on 0,851 m center-to-center<br />

spacing. Rehabilitation work undertaken<br />

in 2000 through 2002 included<br />

deck replacement, new operating<br />

machinery, machinery house restoration,<br />

bridge rail retrofit and some<br />

repairs to the timber approach spans.<br />

The FRP deck was a 178 mm thick<br />

product, which was delivered in 2,134<br />

× 6,199 m sections. Holes with 102 mm<br />

diameter were sawed through the top<br />

and bottom skins of the FRP decking<br />

on 610 mm spacing above the stringers,<br />

and foam dams were placed in the<br />

FRP panels to create 203 × 161 mm<br />

voids around each future shear stud<br />

location (refer to Figs. 1 and 2). The<br />

existing timber deck was removed,<br />

and the top flange of the stringers was<br />

cleaned by power tool cleaning according<br />

to Society for Protective Coatings<br />

(SSPC) SP11 standards. 1 Narrow neoprene<br />

strips were placed to support<br />

the FRP deck panels atop the stringers<br />

Welded shear stud<br />

Portland cement grout<br />

FRP panel<br />

Neoprene strip &<br />

Light gauge steel angle<br />

Steel stringer<br />

Portland cement<br />

grout pad<br />

Section view<br />

looking parallel to roadway centerline<br />

Fig. 1: Longitudinal detail view of FRP<br />

deck installation at shear stud location<br />

Welded shear stud<br />

Portland cement grout<br />

FRP panel<br />

Steel stringer<br />

Foam dam<br />

Portland cement<br />

grout pad<br />

Section view<br />

looking transverse to roadway centerline<br />

Fig. 2: Transverse detail view of FRP deck<br />

installation at shear stud location<br />

418 Technical Report Structural Engineering International 4/<strong>2010</strong>


until the 25 mm cementitious grout<br />

pads were poured. Each FRP deck<br />

panel was placed on the stringers by<br />

a crane as shown in Fig. 3, and then<br />

jacked to seat the tongue and groove<br />

connection between panels, which was<br />

coated with adhesive. Wet-layup FRP<br />

doublers were added over each field<br />

splice. Light gauge steel angles were fastened<br />

to the bottom of the FRP panels<br />

along the sides of each stringer. Steel<br />

shear studs with 22 mm diameter were<br />

welded to the stringers through the<br />

holes in the deck panels. Cementitious<br />

grout filled the gap between the stringers<br />

and the FRP, and filled the pockets<br />

around each shear stud. Cells of the<br />

FRP panels near the joint armor at<br />

each end of the deck were also filled<br />

with cementitious grout.<br />

Following installation of the FRP<br />

decking, a 50 mm thick layer of asphalt<br />

concrete wearing surface was placed<br />

on the bridge. On 25 June 2001 during<br />

installation of the new operating<br />

machinery, the bascule leaf was left in<br />

the “open” position for approximately<br />

5 h. The result was cracking of the<br />

wearing surface, followed by large sections<br />

of asphalt sliding to the bottom<br />

of the open span. 2 Figure 4 shows the<br />

bridge with the bascule leaf in “open”<br />

position after the asphalt had slid-off.<br />

Fig. 3: FRP deck panel being placed on Lewis and Clark River Bridge<br />

Engineering studies and some laboratory<br />

tests 3 were conducted following<br />

this failure. The testing measured<br />

the bond shear properties of asphalt<br />

and identified the fact that asphalt is<br />

a viscoelastic material. When the bascule<br />

leaf was lifted, gravitational force<br />

produced a constant shear force acting<br />

on the asphalt and its bond, producing<br />

viscoelastic strains and eventual<br />

failure. Due to this behavior, it was<br />

concluded that asphalt is not suitable<br />

for an FRP deck on a bascule-type<br />

movable bridge. Recommendations<br />

were made to remove the asphalt and<br />

the “tack” coat, and replace it with a<br />

50 mm thick layer of epoxy polymer<br />

concrete. The old wearing surface was<br />

removed by the use of an excavator,<br />

solvent cleaning, and pressure washing.<br />

The exposed surface was dried by<br />

the use of “weed burner” torches. A<br />

primer was applied to the deck using<br />

paint rollers, and the epoxy binder was<br />

pre-mixed with aggregate, placed on<br />

the deck, and screeded. A topping coat<br />

was applied by the broom and seed<br />

method, where the deck was flooded<br />

with the polymer and the aggregate<br />

was broadcast into the curing polymer<br />

to provide skid resistance.<br />

Cracks began to appear in the epoxy<br />

polymer concrete wearing surface by<br />

May 2002, typically at the locations of<br />

shop and field splices in the FRP deck. 4<br />

Later ODOT tensile tests of polymer<br />

overlay materials 5 demonstrated that<br />

the epoxy polymer concrete exhibits<br />

brittle behavior at temperatures up<br />

to 38°C, making it unable to accommodate<br />

the strains concentrated at<br />

the splice locations when exposed to<br />

normal operating temperatures. These<br />

cracks have continued to grow at a<br />

moderate rate and replacement of the<br />

wearing surface will be required in the<br />

future.<br />

Old Young’s Bay Bridge<br />

Fig. 4: Asphalt concrete wearing surface failure on Lewis and Clark River Bridge<br />

The Old Young’s Bay Bridge, shown<br />

in Fig. 5, is a double-leaf bascule-type<br />

movable bridge built in 1921 that carries<br />

two traffic lanes on a 6,325 m wide<br />

roadway supported by steel stringers<br />

on 1,524 m center-to-center spacing.<br />

Rehabilitation work undertaken<br />

in 2000 through 2002 included deck<br />

replacement, operator’s house restoration,<br />

bridge rail retrofit, and some<br />

repairs to the timber approach spans.<br />

The FRP deck was again a 178 mm<br />

thick product installed using the same<br />

details as on the Lewis and Clark River<br />

Bridge. Following installation of the<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 419


Siuslaw River Bridge<br />

Fig. 5: Historic view of Old Young’s Bay Bridge in the “open” position<br />

FRP decking, a 19 mm thick epoxy<br />

polymer concrete wearing surface was<br />

placed by the broom and seed method.<br />

Cracks began to appear on the epoxy<br />

polymer concrete wearing surface<br />

by May 2002 similar to the cracking<br />

observed on the Lewis and Clark<br />

River Bridge. These cracks continued<br />

growing at an aggressive rate, 4 probably<br />

exacerbated by increased deflections<br />

due to defective shop splices. An<br />

advanced example of this cracking<br />

is shown in Fig. 6. The cementitious<br />

grout around the shear studs failed<br />

as shown in Fig. 7. The cementitious<br />

grout between the FRP panels and<br />

the stringers also failed, starting at<br />

the joint armor and propagating along<br />

the stringers. Due to these failures<br />

and consequent FRP panel damage,<br />

this deck was replaced in June <strong>2010</strong><br />

with a steel grid deck at the request<br />

of ODOT district management. These<br />

failures could have been avoided by<br />

using ductile materials for the wearing<br />

surface and using strong but flexible<br />

materials for the attachment details<br />

to accommodate the deformations<br />

experienced between the flexible steel<br />

floor system and the more flexible<br />

FRP deck panels.<br />

Fig. 6: Cracking and spalling of epoxy polymer concrete wearing surface on the Old<br />

Young’s Bay Bridge<br />

The Siuslaw River Bridge is a doubleleaf<br />

bascule-type movable bridge built<br />

in 1936, carrying two traffic lanes on<br />

an 8,230 m wide roadway, which is<br />

supported by steel stringers on 0,914 m<br />

center-to-center spacing. Retrofit work<br />

undertaken in 2005 included deck<br />

replacement, recycled plastic lumber<br />

sidewalk decking, and new reinforced<br />

plastic walers on the fender system.<br />

The steel grid deck had been supporting<br />

the sidewalks, and the FRP deck<br />

retrofit required installation of new<br />

supports made from rectangular steel<br />

tubing, spanning between sidewalk<br />

brackets to keep the loads off the FRP<br />

decking. Stringer splices were modified<br />

by moving the splice plates to the bottom<br />

of the stringer flange and installing<br />

with countersunk fasteners. The<br />

FRP deck was a 127 mm thick product<br />

which was delivered in 2,438 × 4,089 m<br />

sections. The existing steel grid deck<br />

was removed, and the top flange of the<br />

stringers was repaired and cleaned by<br />

needle guns to SSPC SP11 standards. 1<br />

Neoprene pads with 3 mm thickness<br />

were placed on each stringer. Each<br />

FRP deck panel was placed on the<br />

stringers by a crane, and then jacked to<br />

seat the tongue and groove connection<br />

between panels, which was coated with<br />

adhesive. Wet-layup FRP doublers<br />

were added over each field splice. The<br />

FRP deck panels were secured with 16<br />

mm diameter structural blind fasteners<br />

through existing holes in the stringer<br />

flanges and newly drilled matching<br />

holes in the FRP. This installation is<br />

shown in Fig. 8. Cells of the FRP panels<br />

near the joint armor at each end of<br />

each deck were filled with urethane.<br />

The deck was installed in two stages<br />

to meet traffic needs, with the FRP<br />

panels spliced at roadway centerline.<br />

The bottom skins of the FRP panels<br />

were fastened securely to the flange<br />

of a stringer located at the centerline.<br />

The top skins of the FRP panels<br />

were fastened together using a 2,7<br />

mm thick 304 stainless steel plates,<br />

abrasive blasted and installed using a<br />

polyurethane adhesive sealant and 6,4<br />

mm diameter stainless steel structural<br />

blind fasteners.<br />

Owing to the problems with the polymer<br />

concrete wearing surfaces on the<br />

Astoria bridges, further laboratory<br />

tension testing 5 was performed with<br />

a variety of polymer concrete wearing<br />

surface materials cast into dogbone-shaped<br />

specimens. This testing<br />

revealed that the epoxy polymer<br />

420 Technical Report Structural Engineering International 4/<strong>2010</strong>


e prevented by saw-cutting to break<br />

the bond and filling with sealant. There<br />

has been some loss of aggregate from<br />

the wearing surface, which may be<br />

attributed in some part to the use of<br />

unclean, dusty aggregate that did not<br />

adhere properly to the urethane.<br />

Conclusion<br />

Fig. 7: Grout pocket distress on Old Young’s Bay Bridge, following loss of grout pad<br />

Fig. 8: Attachment details of FRP deck installation on the Siuslaw River Bridge<br />

concrete wearing surface is ductile<br />

only at temperatures over 38°C, but<br />

urethane polymer concrete remains<br />

ductile at temperatures as low as<br />

−9°C. Urethane polymer concrete was<br />

selected for the Siuslaw River Bridge<br />

on the basis of this testing. The FRP<br />

deck was abrasive blasted and then reabrasive<br />

blasted at the request of the<br />

urethane supplier. The deck was then<br />

cleaned using acetone solvent and<br />

a 13 mm thick wearing surface was<br />

placed by the broom and seed method<br />

in two lifts. The completed installation<br />

is shown in Fig. 9.<br />

To date, this installation made with a<br />

ductile wearing surface and strong but<br />

flexible attachment details shows none<br />

of the distress experienced on the two<br />

bridges at Astoria. There are no cracks<br />

in the wearing surface over the shop<br />

and field splices in the FRP deck. At<br />

the ends of the deck, the edges of the<br />

FRP panels deflect and the joint armor<br />

does not, causing a crack that can<br />

Through these experiences, ODOT has<br />

learned that FRP decks are ideal for<br />

replacing decks on movable bridges,<br />

but only if appropriate details are<br />

used for attachment and for wearing<br />

surfaces. The weight of FRP decking<br />

and wearing surface can minimize the<br />

need for rebalancing the bridge, and<br />

the polymer concrete wearing surface<br />

is safer for vehicles including motorcycles<br />

and bicycles. This wearing surface<br />

is also much quieter than steel grid<br />

decking.<br />

It is important to note that the two<br />

bridges at Astoria did not experience<br />

FRP deck failure. They experienced<br />

failure of attachment details and<br />

wearing surfaces that were unable to<br />

accommodate the significant deflection<br />

of FRP.<br />

In addition, FRP decks have potential<br />

for rapid construction and staging<br />

applications as demonstrated by the<br />

two-stage installation accomplished on<br />

the Siuslaw River Bridge.<br />

The following guidelines are recommended<br />

for FRP deck retrofit installations<br />

for moveable bridges:<br />

1. Strong but flexible attachment<br />

details should be used and brittle<br />

materials avoided. ODOT has had<br />

good experience with structural<br />

blind fasteners with neoprene pads<br />

between the stringers and the FRP<br />

panels.<br />

2. The number of holes cut into the<br />

top and bottom skins of the FRP<br />

panels should be minimized.<br />

3. Polymer concrete wearing surface<br />

having high degree of ductility at<br />

expected operating temperatures<br />

should be used.<br />

4. Clean, dry aggregate should be<br />

used in any polymer concrete<br />

wearing surface.<br />

5. Asphalt concrete on bascule-type<br />

movable bridges should not be<br />

used.<br />

6. Adequate drain holes should be<br />

provided to prevent accumulation<br />

of water in FRP panels.<br />

7. When replacing steel grid decking,<br />

the likelihood that the stringers<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 421


quality control programs should be<br />

specified.<br />

10. A saw-cut and flexible sealant<br />

between the FRP deck panels<br />

and rigid joint armor should be<br />

provided.<br />

References<br />

[1] Surface Preparation Standard No. 11,<br />

Power Tool Cleaning to Bare Metal. SSPC: The<br />

Society for Protective Coatings. Pittsburgh,<br />

PA, September 1, 2000 and November 1, 2004<br />

revisions.<br />

[2] Nelson J. Asphalt Crashes Down From<br />

Bridge. The Oregonian. Portland, Oregon, June<br />

27, 2001.<br />

[3] Lovejoy SC, Nelson SD, Patterson BM.<br />

Causes and Solutions to the Failure of the Wear<br />

Surface on the Lewis and Clark River Bridge.<br />

Unpublished report. Oregon Department of<br />

Transportation, November 21, 2001.<br />

Fig. 9: View of completed FRP deck installation on the Siuslaw River Bridge<br />

will have wear damage and may<br />

require replacement should be<br />

considered.<br />

8. Stringer splices with plates on top<br />

of the top flange should be modified<br />

to a configuration with countersunk<br />

fasteners and splice plates<br />

on the underside of the top flange.<br />

9. FRP deck products made by<br />

manu facturers that have effective<br />

[4] Lovejoy SC. Summary of Cracking in the<br />

Wearing Surface of the FRP Decks on Bridges<br />

0711 (Lewis and Clark River) and 0330 (Old<br />

Young’s Bay). Unpublished report. Oregon<br />

Department of Transportation, October 20, 2004.<br />

[5] Barquist G, Lovejoy SC, Nelson SD, Soltesz<br />

S. Evaluation of Wearing Surface Materials<br />

for FRP Bridge Decks. Oregon Department<br />

of Transportation and Federal Highway<br />

Administration, July, 2005.<br />

Make Your Client<br />

Remember You<br />

through out the Year..<br />

Gift a <strong>SEI</strong> Subscription<br />

Contact: iabse@iabse.org, Subject: Subscription Gift<br />

CHF 220/Year for 4 issues.<br />

422 Technical Report Structural Engineering International 4/<strong>2010</strong>


Glass Fiber Reinforced Polymer Strengthening<br />

and Evaluation of Railroad Bridge Members<br />

GangaRao V.S. Hota, Prof., Ph.D., P.E., Director; P.V. Vijay, Research Prof., PhD, PE; Dept. of Civil and Environmental<br />

Engineering West Virginia University, Morgantown, USA and Reza S. Abhari, MSCE, PE, Structural Eng., Alpha Corporation; USA.<br />

Contact: p.vijay@mail.wvu.edu<br />

Abstract<br />

In this work, damaged timber railroad<br />

bridge stringers and piles were<br />

rehabilitated with glass fiber reinforced<br />

polymer (GFRP) composites,<br />

and tested. Four timber stringers (152<br />

× 203 × 3560 mm) removed from the<br />

field were rehabilitated with GFRP<br />

spray lay-up and GFRP wrap vacuum<br />

bagging methods. GFRP strengthening<br />

increased the shear moduli of<br />

the two stringers by 41 and 267%.<br />

Rehabilitation and load testing were<br />

also performed on an open-deck-timber<br />

railroad bridge built during the<br />

early 1900s on the South Branch Valley<br />

Railroad (SBVR) owned by the West<br />

Virginia Department of Transportation<br />

(WVDOT) in Moorefield, WV,<br />

USA. Specifically, field rehabilitation<br />

involved repairing piles using GFRP<br />

composite wraps and phenolic formaldehyde<br />

adhesives. Static and dynamic<br />

tests using a 80 ton locomotive showed<br />

that the rehabilitated piles and pile cap<br />

showed a 43 and 46% strain reduction,<br />

respectively. Dynamic load amplification<br />

factor was noted to be almost<br />

close to a speed of 24 km/h.<br />

Keywords: strengthening; GFRP; composites;<br />

rehabilitation; timber piles;<br />

timber stringers.<br />

Introduction<br />

In the USA, timber was the primary<br />

railroad bridge construction material<br />

until the early 20th century. The use<br />

of lumber began to slowly decline in<br />

the early 20th century as other materials<br />

such as steel and concrete replaced<br />

wood for bridge construction. The US<br />

railroads have more than 2400 km of<br />

timber bridges and trestles in service<br />

today, which typically provide more<br />

than 50 years of acceptable service<br />

when properly designed, constructed,<br />

and maintained. 1 With the recent<br />

increases in railway car axle loads<br />

and long-term exposure to the environment,<br />

many of these bridges have<br />

reached the end of their service life or<br />

their load carrying capacity has lowered<br />

significantly because of material<br />

aging. This work describes the use of<br />

GFRP for rehabilitating in-service<br />

superstructure and substructure railroad<br />

timber bridge members.<br />

Objectives and Scope<br />

The work presented herein focuses on<br />

developing safer, faster, and most practical<br />

methods that allow in situ rehabilitation<br />

of timber railroad bridges<br />

without interrupting rail traffic. The<br />

objective of this research work is to<br />

evaluate the application of GFRP<br />

composite spray and wraps as a viable<br />

rehabilitation alternative for stringers<br />

and piles on in-service timber<br />

bridges. In this work, damaged timber<br />

bridge stringers and piles selected<br />

by the South Branch Valley Railroad<br />

(SBVR) in Moorefield, WV, USA were<br />

repaired and rehabilitated with the use<br />

of GFRP composites. The work consisted<br />

of: (a) determining bending and<br />

shear properties of damaged timber<br />

stringers removed from field-service,<br />

and performing analytical comparisons<br />

prior to and after repair through<br />

sprayed GFRP (SGFRP) and vacuum<br />

bagging methods, and (b) rehabilitating<br />

and evaluating timber railroad<br />

bridge piles under static and dynamic<br />

loads (up to 24 km/h).<br />

Stringer Rehabilitation Scheme<br />

A total of four 50+ year-old creosotetreated<br />

Douglas-fir stringers were<br />

acquired from the SBVR-WVDOT.<br />

The four stringers were deemed deficient<br />

due to significant checking and<br />

splitting along their length (Fig. 1). All<br />

four specimens (152 × 203 × 3560 mm)<br />

were load tested up to 26,7 kN before<br />

repair to determine flexural rigidity<br />

and shear modulus. Stringers were surface<br />

prepared by removing dirt and<br />

debris, by filling of void or splits with<br />

fillers (resin and saw dust mix), and<br />

edge smoothening to reduce stress<br />

concentrations. After sanding, primer<br />

coat (phenolic resin) was applied and<br />

cured. Retrofitting was done using the<br />

most compatible primer/resin combination<br />

using phenolic formaldehyde.<br />

Vacuum Bagging and<br />

GFRP Spray Methods<br />

Vacuum bagging creates clamping<br />

forces (127–152 mm Hg vacuum)<br />

to hold the resin-coated fiber/fabric<br />

components in place and removes the<br />

trapped air. Fabric pieces were cut<br />

to cover three sides of the specimen.<br />

The vacuum bagging method used a<br />

sealed plastic bag covering the lay-up<br />

area including a breather layer, and a<br />

vacuum pump to remove the air and<br />

apply uniform surface pressure during<br />

resin curing (Fig. 1). For the SGFRP<br />

method, two separate compressed air<br />

spray guns were used, in lieu of an<br />

Vacuum bag<br />

covering the<br />

wrapped area<br />

Vacuum pump<br />

(a) (b) (c)<br />

Fig. 1: Rehabilitation of timber stringers: (a) field-removed stringers; (b) vacuum bagging of wrapped beam; (c) spraying fiber on the<br />

wet surface<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 423


integrated sprayer, one for spraying<br />

the resin and another for spraying the<br />

fiber. Chopped glass fibers, 12,7 mm<br />

in length, were sprayed on the primer<br />

coated wood surface in six cycles with<br />

a final SGFRP layer thickness measuring<br />

15 to 18 mm. Unlike the wrapping<br />

method that requires a smooth surface,<br />

spraying technique can easily be<br />

applied to any type of surface.<br />

Stringer Tests/Analysis<br />

After rehabilitation, stringers were<br />

tested under four-point bending with a<br />

shear span (a/h) ratio of 3,4. Stringers<br />

were instrumented with strain gages on<br />

the surface of the GFRP wraps at midspan<br />

on the tension side and rosette<br />

gages at the neutral axis (to evaluate<br />

shear and principal strains). Load versus<br />

deflection and shear strain were plotted<br />

and the modes of failure were identified.<br />

Based on the rule of mixtures and<br />

burn-off tests, the average elastic moduli<br />

for the GFRP wraps and SGFRP specimens<br />

were found to be 18 340 and 8963<br />

N/mm 2 , respectively. The lower modulus<br />

of the SGFRP is attributed to the lower<br />

density of the sprayed wraps.<br />

The results in Table 1 indicate that<br />

repairing damaged timber stringers<br />

using GFRP or SGFRP is effective in<br />

improving their flexural rigidity and<br />

shear moduli. Repairing shear zones of<br />

a stringer with GFRP composites had a<br />

significant impact on the shear modulus<br />

and deflection response. As the tested<br />

stringers were in poor condition with<br />

large voids, they were filled with strong<br />

filler material. Specimen #2 had wide<br />

longitudinal splitting, therefore the initial<br />

EI value was low, but it had more<br />

than tripled after repair. Specimens<br />

#3 and #4 containing less voids had a<br />

19,5 and 15% increase in the EI value,<br />

respectively. For specimen #3 repaired<br />

with four layers of GFRP wrap with<br />

vacuum bagging, the maximum strain<br />

measured at the extreme tension fibers<br />

at mid-span was 2488 µε compared to<br />

the maximum shear strain measured at<br />

the neutral axis of the specimen, which<br />

was 4446 µε.<br />

Equations (1) and (2) were used to<br />

calculate the parameters shown in<br />

Table 1 for beams prior to and after<br />

rehabilitation with FRP using SGFRP<br />

(beams #2 and #4) and vacuum bagging<br />

(beam #3) techniques. 2 The failure<br />

of a wooden beam in shear occurs<br />

when the maximum shear stress, s max ,<br />

becomes equal to the shear strength<br />

of wood, s d , as given by Eq. (1). 2 The<br />

flexural rigidity (EI exp ) of the repaired<br />

specimens is computed from deflection<br />

equation (2) with bending and<br />

shear terms (Δ = Δ bending + Δ shear ).<br />

v<br />

bh<br />

EI<br />

d<br />

exp<br />

<br />

n h h<br />

+ <br />

<br />

h<br />

+ 2<br />

<br />

<br />

1<br />

frp<br />

1<br />

h<br />

frp<br />

<br />

2<br />

frp <br />

<br />

frp <br />

<br />

<br />

<br />

= <br />

<br />

3 frp <br />

+<br />

frp<br />

<br />

1 n h = 2<br />

h 3 r (1)<br />

<br />

<br />

<br />

<br />

<br />

a P 2 2<br />

( L – a )<br />

<br />

<br />

<br />

3 4<br />

<br />

=<br />

a P<br />

<br />

48 1 –<br />

GA <br />

<br />

<br />

<br />

2 <br />

<br />

<br />

(2)<br />

Where V = shear force; s d = shear<br />

strength of wood in horizontal shear;<br />

b = width of the beam; I = moment of<br />

inertia of the section; h frp = height of<br />

the FRP section; h = height of the beam<br />

section; n = modular ratio = E SGFRP/<br />

E wood ; r frp = FRP area fraction = 2th frp /<br />

bh; t = thickness of the FRP composite;<br />

r = ratio of shear capacity of FRPreinforced<br />

section to shear capacity of<br />

unreinforced section (effectiveness);<br />

L = span; k = shape factor; a = shear<br />

span; and (P/D) is the initial slope of<br />

load versus deflection curve.<br />

In Situ Bridge Pile<br />

Rehabilitation<br />

Overview<br />

In situ pile rehabilitation was carried<br />

out on SBVR-WVDOT track containing<br />

several timber bridges constructed<br />

in the early 1900s. Bridge No. 574 was<br />

approximately 14 m long with four piles<br />

that were highly deteriorated and/or<br />

damaged due to flooding (Fig. 2). Pile<br />

#1 was heart rotted (i.e. damaged core)<br />

and only about 20% of the cross sectional<br />

area was functional, and piles #2,<br />

#3, and #4 were heart rotted with 45%<br />

effectiveness. Instead of replacement,<br />

the badly damaged piles were rehabilitated<br />

to enhance their durability and<br />

load-bearing capacity. Phenolic-based<br />

adhesive with E-glass FRP composite<br />

wrap (0,4 kg/m 2 density with 0,2 kg/m 2<br />

in both 0 and 90 directions) was used<br />

owing to its cost advantage, chemical<br />

compatibility, excellent bonding, and<br />

electrical non-conductivity 3 .<br />

Rehabilitation Details<br />

Cofferdams were built around each<br />

submerged pile using metal sheets.<br />

Piles were pressure washed, dried,<br />

and sanded to obtain a smooth surface.<br />

Phenolic-based adhesive filler<br />

with 20% sawdust was used to fill the<br />

voids, cracks, and partially missing core<br />

of the core-damaged piles. The wrapping<br />

areas were pre-coated and cured<br />

for 6 h at ambient temperature (27°C)<br />

with phenolic-based primer. The precoated<br />

piles were wrapped to a length<br />

of 888 mm with two 482 mm resin wet<br />

GFRP fabrics having 76 mm overlap.<br />

A UV protection layer of resin was<br />

applied to the wrapped areas after<br />

rehabilitation.<br />

Field Testing, Results, Analysis,<br />

and Discussions<br />

Field testing was carried out using an<br />

80 ton locomotive prior to and after<br />

rehabilitation with different axle positions<br />

on the bridge for assessing the<br />

strains and stresses in pile caps and<br />

Specimen/<br />

(repair scheme)<br />

Peak<br />

load (kN)<br />

s max , Shear<br />

strength<br />

(N/mm 2 )<br />

Max. shear<br />

strain to failure<br />

(×10 −6 )<br />

EI, Flexural rigidity (N/mm 2 ) G, Shear modulus (N/mm 2 )<br />

Control<br />

(×10 6 )<br />

Repaired<br />

(×10 6 )<br />

Increase<br />

(%)<br />

Control Repaired Increase<br />

(%)<br />

#2 (SGFRP) 161,5 0,53 2937 2,85 8,7 305 — 216,2 —<br />

#3 (wrapping) 206,6 1,55 4446 11,1 13,3 19,5 295,3 793,6 267<br />

#4 (SGFRP) 209,3 1,28 3379 10,1 11,6 15 270,7 383,6 41<br />

Specimen #1 is control specimen and not repaired. Shear strains on specimen #2 without wrap are not considered.<br />

Table 1: Stringer properties (EI and G) before and after repair<br />

424 Technical Report Structural Engineering International 4/<strong>2010</strong>


Wrapped<br />

height<br />

888 mm<br />

1219<br />

1981<br />

1067<br />

1524<br />

1676<br />

All dimensions<br />

in mm<br />

Average<br />

water<br />

level<br />

Pile 3<br />

Pile 2<br />

Pile 4 Pile 1<br />

305<br />

610<br />

508<br />

1575<br />

457<br />

3658<br />

Strain<br />

gage<br />

D = 305 D = 610 D = 610 D = 305<br />

(a) (b) (c)<br />

Fig. 2: Pile rehabilitation: (a) water submerged piles; (b) pile repaired with GFRP; (c) pile configuration and strain gage locations<br />

Loading<br />

Micro-strains<br />

Center of rear axle Before rehabilitation — −243 −144 152 230<br />

on pile bent<br />

After rehabilitation −185 −138 −127 108 129<br />

Reduction (%) 43 12 29 44<br />

24 km/h Before rehabilitation — −364 −150 278 253<br />

After rehabilitation −376 −229 −134 119 146<br />

Reduction (%) 37 11 57 42<br />

Negative (−) strain–compression; positive (+) strain–tension.<br />

Pile<br />

4<br />

Pile<br />

3<br />

Table 2: Results for different load cases before and after rehabilitation<br />

piles. The piles were not symmetrically<br />

placed when the bridge was built<br />

and this can be seen from the dimensions<br />

in Fig. 2. After rehabilitation, the<br />

strains induced in the timber piles and<br />

pile caps under different load conditions<br />

were reduced up to 43 and 46%,<br />

respectively (Table 2), with improved<br />

load distribution. Pile #2 was not rehabilitated<br />

and was considered as a control<br />

pile. It was also noted that strain<br />

values for pile #1 were in tension,<br />

instead of being in compression, which<br />

Pile<br />

2<br />

Pile<br />

1<br />

Pile<br />

Cap<br />

is attributed to local stress patterns and<br />

pile buckling at the top section and the<br />

displacement of the whole pile from<br />

its original axis. Dynamic load testing<br />

up to the SBVR-DOT recommended<br />

maximum speed on the bridge showed<br />

that the strains were similar to static<br />

values with a dynamic load amplification<br />

(impact) factor of 1 up to a speed<br />

of 24 km/h (Fig. 3). Dynamic load testing<br />

also showed a strain reduction in<br />

the FRP rehabilitated piles #1 and #3,<br />

and showed no change in the strain<br />

value of control pile #2, which was not<br />

rehabilitated (Fig. 3).<br />

Conclusion<br />

Damaged stringers strengthened with<br />

GFRP and tested under controlled<br />

laboratory conditions showed a 267<br />

and 41% improvement in the shear<br />

modulus when repaired with wrapping<br />

method and SGFRP method, respectively.<br />

The 50+ year-old in situ rehabilitated<br />

timber railroad bridge subjected<br />

to static and dynamic loads using an<br />

80 ton locomotive, revealed a reduction<br />

in pile cap strains by approximately<br />

44% and in the piles by approximately<br />

57%. The dynamic effect was negligible<br />

at the speed ranges up to 24 km/h<br />

in most of the piles. GFRP composite<br />

fabric in combination with phenolic<br />

formaldehyde adhesives was found to<br />

perform well with excellent bonding<br />

under harsh environments for GFRP<br />

wrapped timber piles.<br />

Acknowledgement<br />

Funding provided by Federal Railroad<br />

Authority (FRA) and West Virginia<br />

300<br />

200<br />

Pile 1: before rehab<br />

After rehab<br />

100<br />

Micro strains<br />

−100<br />

−200<br />

0<br />

10 11 12 13 14 15<br />

Pile 2: before rehab<br />

After rehab<br />

(a)<br />

−400<br />

(b)<br />

Pile 3: before rehab<br />

After rehab<br />

Fig. 3: Dynamic load testing: (a) 80 ton locomotive; (b) time versus strain data for piles at 24 km/h<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 425<br />

−300


Division of Highways (WVDOH),<br />

USA for this work is acknowledged.<br />

References<br />

[1] Ritter M. Timber Bridges. Construction,<br />

Inspection and Maintenance. United States<br />

Department of Agriculture Forest Service, 1990.<br />

[2] Triantafillou T. Shear reinforcement of wood<br />

using FRP materials. J. Mater. Civil Eng. ASCE,<br />

9(2), 1997.<br />

[3] Abhari RS. Rehabilitation of Timber Railroad<br />

Bridges using GFRP Composites, MS Thesis,<br />

West Virginia University, 2007.<br />

<strong>SEI</strong> Data Block<br />

Owner:<br />

South Branch Valley Railroad,<br />

Authority, West Virgin ia (WV)<br />

Department of Transportation (USA)<br />

Designer:<br />

Constructed Facilities Center, West<br />

Virginia University, Morgantown, WV.<br />

Main contractor:<br />

Constructed Facilities Center,<br />

West Virginia University,<br />

Morgantown, WV.<br />

FRP fabricator:<br />

Constructed Facilities Center,<br />

West Virginia University,<br />

Morgantown, WV.<br />

FRP (t): 0, 1<br />

Span lengths (m): 3560 mm<br />

(stringers) and 14000 mm (bridges)<br />

Rehabilitation cost (USD million):<br />

Less than 10,000<br />

Service date: 2001/2004 (Different<br />

times of rehabilitation)<br />

426 Technical Report Structural Engineering International 4/<strong>2010</strong>


Design of the St Austell Fibre-Reinforced Polymer<br />

Footbridge, UK<br />

Jonathan Shave, Principal Eng., Parsons Brinckerhoff, Queen Victoria House, Redland Hill, Bristol, UK; Steve Denton, Engineering<br />

Director, Parsons Brinckerhoff, UK; Ian Frostick, Territory Civil Eng., Network Rail, UK. Contact: shavej@pbworld.com<br />

Abstract<br />

The St Austell Footbridge is a lightweight<br />

glass fibre-reinforced polymer<br />

(FRP) structure comprising pultruded<br />

and moulded elements, which crosses<br />

a railway line in St Austell, UK. When<br />

it was installed in October 2007, it was<br />

the first structure on the UK rail network<br />

to be entirely constructed from<br />

FRP materials. The bridge structure<br />

was designed to satisfy the aesthetic<br />

and environmental requirements of<br />

the client. Through rapid installation<br />

and the minimisation of any maintenance<br />

requirements, it has delivered<br />

economic, operational and sustainability<br />

benefits.<br />

This paper presents some of the design<br />

issues encountered in developing the<br />

innovative structure. These include the<br />

development of techniques and design<br />

philosophies to enhance the structure’s<br />

robustness through careful consideration<br />

of the potential behaviour of<br />

the structure and its components and<br />

connections up to the ultimate limit<br />

state. It also describes how the design<br />

was developed to ensure that vibrations<br />

caused by trains or pedestrians<br />

would be adequately controlled at<br />

the serviceability limit state. The findings<br />

of the research regarding behaviour<br />

of lightweight structures subject<br />

to aerodynamic impulses associated<br />

with passing trains are presented, and<br />

the corresponding effect on the design<br />

development is described.<br />

Also described in this paper are the<br />

regimes of testing and monitoring<br />

that were developed to ensure that<br />

the structure and its components were<br />

behaving as expected.<br />

Keywords: fibre-reinforced polymer;<br />

advanced composites; footbridge;<br />

design; robustness; vibration; buffeting;<br />

innovation.<br />

the town of St Austell, Cornwall, UK.<br />

The footbridge over the railway track<br />

gives pedestrians access to various<br />

amenities including a doctors’ surgery<br />

and a leisure centre.<br />

The old footbridge comprised three<br />

simply supported spans of approximately<br />

5, 14 and 6 m, respectively, and<br />

was supported by masonry piers and<br />

abutments. It was proposed to replace<br />

the superstructure with a new fibrereinforced<br />

polymer (FRP) structure<br />

and adapt the existing substructure for<br />

the new bridge (Fig. 1).<br />

Design Philosophy<br />

Design Concept<br />

The design concept for the new footbridge<br />

was developed on the basis of<br />

the Advanced Composite Construction<br />

System (ACCS), a modular system<br />

of pultruded structural panels. The<br />

system comprises cellular panels and<br />

boxes that are connected using adhesive<br />

joints and additionally secured<br />

with inserted mechanical connectors.<br />

This arrangement allows an anisotropic<br />

thick shell to be built up from standard<br />

modules in a variety of possible configurations.<br />

The designers recognised<br />

that it would be possible to use this<br />

system innovatively and efficiently in<br />

a U-frame configuration, as illustrated<br />

in Fig. 2.<br />

This approach minimised the construction<br />

depth from footway level to soffit<br />

level, and allowed the parapet walls to<br />

be used as structural members, giving<br />

the bridge sufficient stiffness to satisfy<br />

deflection requirements. Transverse<br />

frames were provided at regular intervals<br />

to resist distortional buckling of<br />

the cross section.<br />

Robustness<br />

As with any structural design, it was<br />

important to consider the robustness of<br />

the structure and ensure that progressive<br />

non-ductile failure modes would<br />

be avoided. The method of achieving<br />

this objective in the design of the FRP<br />

structure differed from conventional<br />

structures, as it was not appropriate to<br />

rely on any ductile behaviour of the<br />

FRP material or the adhesive joints.<br />

The high strength-to-stiffness ratio and<br />

low mass of the FRP material had the<br />

effect that the design was governed<br />

by serviceability limit states (such as<br />

deflection and vibration) rather than<br />

ultimate limit states. The factor of<br />

Introduction<br />

The St Austell Footbridge was required<br />

to replace a corroded wrought iron<br />

and cast iron structure over the<br />

Paddington–Penzance railway line in<br />

Fig. 1: St Austell Footbridge in service<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 427


Longitudinal elements comprising<br />

ACCS panels built up into U-shape<br />

Fig. 2: U-frame concept<br />

Exterior skin<br />

comprising<br />

moulded panels<br />

Transverse frames to provide<br />

lateral stiffness and stability<br />

safety on the ultimate strength of the<br />

footbridge was high, and the theoretical<br />

load required to cause structural<br />

failure would never be encountered<br />

(and even if it were, the structure<br />

would already have become unserviceable<br />

due to excessive deflection).<br />

Notwithstanding this high factor of<br />

safety, there remained the possibility<br />

that an abnormal stress state could be<br />

triggered, for example, due to the unexpected<br />

failure of a joint. This scenario<br />

was considered carefully in the design<br />

of St Austell Footbridge, to avoid the<br />

possibility of a non-ductile progressive<br />

collapse of the structure.<br />

The vulnerability of the joints was initially<br />

minimised in the design by avoiding<br />

any in-span transverse joints and<br />

by careful specification of the adhesive<br />

(with associated testing) to ensure<br />

a sufficiently high factor of safety. In<br />

addition, the joints were given further<br />

robustness through the additional<br />

mechanical connector that effectively<br />

clamps each bonded joint together, significantly<br />

reducing the risk of any Mode<br />

I fracture behaviour. Notwithstanding<br />

these features, it was recognised that<br />

the potential vulnerability of the joints<br />

should be a key factor that had to be<br />

taken into account in the design stage<br />

when considering robustness.<br />

The philosophy was that the structure<br />

should be able to withstand the loss<br />

of any joint without causing ultimate<br />

collapse of the structure. Hence, if the<br />

most heavily loaded joints at the base<br />

of each parapet were to unexpectedly<br />

fail, losing the composite action<br />

between the parapets and the deck,<br />

then the deck section should have sufficient<br />

ultimate strength acting on its<br />

own to resist nominal loading (including<br />

pedestrian loading) without collapse.<br />

In this situation, there would<br />

be an increase in deflection and other<br />

noticeable indicators of distress. This<br />

behaviour is characteristic of a robust<br />

structure, improving safety and allowing<br />

effective monitoring techniques to<br />

be used.<br />

Outer Skin<br />

A further degree of robustness was<br />

achieved by encasing the main pultruded<br />

structural elements in a tough<br />

outer skin, providing additional protection<br />

from damage due to vandalism<br />

or adverse environmental conditions<br />

(such as ultraviolet radiation). Moulded<br />

FRP panels were used on the outside<br />

of the structure. These have the added<br />

advantage of improving the aesthetics<br />

and providing smooth curves to the<br />

structure. The inner faces of the parapets<br />

are protected with tough, replaceable<br />

anti-vandal panels, which prevent<br />

surface damage to the structure and<br />

have an anti-graffiti “deep-crinkle”<br />

finish. The non-slip walking surface<br />

was provided by a tough gritted FRP<br />

plate.<br />

Design for Vibration Serviceability<br />

The use of very light FRP materials<br />

allows the structure to have a low mass<br />

of only 5 t for the central 14 m span.<br />

There are various benefits to having<br />

a very light structure, principally concerned<br />

with installation and transportation,<br />

but the low mass also had some<br />

interesting effects on the vibration serviceability<br />

design, particularly under<br />

the actions of passing trains. This aspect<br />

of the design became the governing criterion,<br />

and required some cutting-edge<br />

research to be carried out in order to<br />

provide sufficient assurance that the<br />

structure would be serviceable.<br />

Two aspects of vibration serviceability<br />

were considered: pedestrian-induced<br />

vibration and train-induced vibration.<br />

Owing to the low mass of the structure,<br />

the design natural frequencies<br />

were generally higher than would be<br />

normal for a conventional footbridge.<br />

The fundamental vertical mode had a<br />

frequency of 12 Hz, and the horizontal<br />

mode had a frequency of 7 Hz, both<br />

of which were high enough for the<br />

structure to generally be deemed adequate<br />

for pedestrian-induced vibration<br />

serviceability. Even allowing for the<br />

unusually low mass of the structure,<br />

the accelerations arising from pedestrian<br />

actions were not expected to be<br />

high, and this proved to be the case.<br />

Vibrations arising from the aerodynamic-buffeting<br />

action of trains passing<br />

under the footbridge were harder<br />

to quantify. It was important to be able<br />

to demonstrate that the design accelerations<br />

would not cause discomfort to<br />

the pedestrians using the structure.<br />

Existing design standards provided only<br />

very limited guidance on the appropriate<br />

loadings to be used to model<br />

train buffeting effects. EN1991-2 1<br />

recommends a loading model, but this<br />

is not appropriate for use in the United<br />

Kingdom, as stated in the accompanying<br />

UK National Annex. 2<br />

Accordingly, it was necessary to derive<br />

a new loading model. An extensive<br />

literature search suggested that there<br />

was very little useful data on train buffeting<br />

vibrations, and so it was necessary<br />

to carry out new research based<br />

on actual measurements to derive an<br />

appropriate loading model.<br />

The designers, worked with Sheffield<br />

University experts to take acceleration<br />

measurements of a temporary footbridge<br />

over the railway at Goring, UK, which<br />

was exposed to considerable vibrations<br />

due to buffeting. These measurements<br />

were then back analysed to derive a proposed<br />

loading model for train buffeting<br />

to be used for the St Austell Footbridge<br />

design. Even using this loading model,<br />

and allowing for a reasonably modest<br />

line speed of 105 km/h at St Austell, the<br />

design of the footbridge was still governed<br />

by the limitation of vibrations due<br />

to train buffeting effects at the serviceability<br />

limit state.<br />

The development of the train buffeting<br />

load model allowed the footbridge<br />

design to be completed in an optimised<br />

and lightweight form, without the need<br />

for additional ballast or dampers to<br />

reduce theoretical vibration levels.<br />

428 Technical Report Structural Engineering International 4/<strong>2010</strong>


Construction<br />

Fabrication of each span was achieved<br />

off site in factory conditions by<br />

building up the cross section from<br />

the ACCS components using adhesive<br />

bonds and mechanical inserts along<br />

the entire length of each span.<br />

Transverse U-frames were incorporated<br />

at regular intervals, with the<br />

horizontal frame members passing<br />

through the holes in the middle layer<br />

of the ACCS box. The internal faces<br />

were clad with the anti-vandal plates<br />

and the gritted base plate.<br />

The curved moulded panels on the<br />

exterior faces were produced by incorporating<br />

internal stiffening ribs to<br />

provide sufficient resistance to wind<br />

loading, and fixed to the structure at<br />

the top and bottom of each parapet<br />

wall. Moulded capping pieces were<br />

fixed to the top of each parapet wall,<br />

placed over the anti-vandal plate and<br />

the exterior moulded panels.<br />

Each bridge span was transported to<br />

the site as a single prefabricated unit,<br />

and was rapidly installed in a single<br />

night, as illustrated in Fig. 3. The spans<br />

were placed onto elastomeric bearings<br />

and held down to the new concrete<br />

padstones using stainless steel fixings<br />

and brackets. After installation of the<br />

spans, final panels were installed at the<br />

pier positions, providing an unbroken<br />

elevation and allowing relative movement<br />

and rotation at the bearings.<br />

Testing and Monitoring<br />

A significant amount of testing was<br />

carried out for this project, at the following<br />

levels:<br />

– Pultrusion testing<br />

– Adhesive testing<br />

– Moulded component testing<br />

– Complete structure testing (static<br />

and dynamic).<br />

The project specification was developed<br />

to include a comprehensive<br />

Fig. 3: Installation of the main span<br />

series of tests to ensure that all the<br />

materials and components had the<br />

correct properties. These included<br />

short-term and long-term stiffness and<br />

strength tests on the ACCS pultrusions,<br />

fire tests, adhesive tests at ambient and<br />

elevated temperatures and testing of<br />

the moulded panels for stiffness and<br />

strength.<br />

In addition, due to the innovative<br />

nature of the structure it was appropriate<br />

to specify testing of the complete<br />

structure under both static and<br />

dynamic loading conditions.<br />

The main span of the structure was<br />

tested under static loading conditions<br />

in the factory by filling it with 10,1 t<br />

of water up to a maximum level of<br />

510 mm, representing a design serviceability<br />

loading of 5 kPa (Fig. 4).<br />

The testing confirmed the following:<br />

– The structure as fabricated was adequate<br />

to carry crowd loading.<br />

– Deflections under serviceability<br />

loading were not excessive and were<br />

consistent with predictions.<br />

– The load–deflection behaviour was<br />

linear.<br />

To confirm the vibration behaviour<br />

of the footbridge, vibration testing<br />

was carried out a week after installation.<br />

Vertical and horizontal shakers<br />

were used in combination with accelerometers<br />

to obtain a full dynamic<br />

blueprint of the structure, including<br />

modeshapes, frequencies, modal<br />

masses and damping coefficients. The<br />

natural frequencies and mode shapes<br />

for the fundamental lateral and vertical<br />

modes will be monitored over the<br />

life of the structure in order to identify<br />

any reductions in stiffness or integrity<br />

of the structure.<br />

The dynamic response of the main span<br />

to pedestrian excitation was confirmed<br />

to be minor, as predicted. The test subjects<br />

reported that the vibrations were<br />

comfortable at all speeds.<br />

Fig. 4: Static load testing with water<br />

The main span was also tested for<br />

vibrations related to train buffeting<br />

loading and were well within acceptable<br />

serviceability limits. The speed of<br />

the trains was measured using a speed<br />

gun, and the accelerations were back<br />

analysed providing additional data for<br />

further improvement and calibration<br />

of the buffeting load model.<br />

Conclusion<br />

The St Austell Footbridge is a highly<br />

innovative structure, and the first all-<br />

FRP structure to be installed on the<br />

UK rail network.<br />

As a successful “pioneer” project, it<br />

paves the way for the increased use of<br />

these lightweight materials across the<br />

transportation sector to provide robust<br />

and aesthetically pleasing structures<br />

that deliver considerable economic,<br />

operational and sustainability benefits.<br />

Its design included the development<br />

of methods to ensure that a sufficient<br />

degree of robustness would be<br />

achieved.<br />

Research was carried out to ensure<br />

that vibration serviceability would be<br />

satisfied, particularly regarding the<br />

development and refinement of loading<br />

models for train buffeting effects.<br />

A significant degree of testing was carried<br />

out on the structure and its components,<br />

with highly successful results.<br />

References<br />

[1] BS EN 1991–2. Eurocode 1: Actions on<br />

Structures. Part 2: Traffic Loads on Bridges, BSi,<br />

October 2003.<br />

[2] National Annex to BS EN 1991–2. Eurocode 1:<br />

Actions on Structures. Part 2: Traffic Loads on<br />

Bridges, BSi, May 2008.<br />

<strong>SEI</strong> Data Block<br />

Owner:<br />

Network Rail, UK<br />

Designer:<br />

Parsons Brinckerhoff<br />

Main contractor:<br />

BAM Nuttall<br />

FRP fabricator:<br />

Pipex Structural Composites<br />

FRP (t): 8<br />

Span lengths (m): 5, 14, 6<br />

Construction cost<br />

(EUR million): 0,4<br />

Service date: October 2007<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 429


Aluminium Structures in Building and Civil Engineering<br />

Applications<br />

Frans Soetens, Prof., Eindhoven University of Technology, Eindhoven, The Netherlands; TNO Built Environment and Geosciences,<br />

Delft, The Netherlands. Contact: fsoetens@tue.nl<br />

Abstract<br />

Structural applications of aluminium are considered in this paper. Although<br />

the discussion is mainly devoted to Europe, the paper also refers, where possible,<br />

to developments in other parts of the world. The problems faced by a<br />

designer in creating an optimum design are described, followed by a brief<br />

review of the research carried out in the past four decades on the structural<br />

behaviour of aluminium and a preview of topics still to be investigated.<br />

A historical overview of standards is then given, starting from the ECCS<br />

Recommendations up to the recently published Eurocode 9. Finally, a number<br />

of structural applications are dealt with, as well as some future directions and<br />

concluding remarks.<br />

Keywords: aluminium; structures; design; research; standards; applications.<br />

Introduction<br />

Successful design and calculation solutions<br />

for aluminium elements and<br />

structures are only possible when<br />

the favourable properties of aluminium—in<br />

particular, its light weight and<br />

corrosion resistance—are taken into<br />

account, and an adequate design solution<br />

is provided for its unfavourable<br />

properties. These properties are unfavourable<br />

in comparison to the behaviour<br />

of steel structures and include<br />

the low Young’s modulus of aluminium,<br />

resulting in higher deformations,<br />

and a higher sensitivity to stability<br />

issues, as well as its vulnerability to<br />

fatigue issues and its lower fire resistance.<br />

Furthermore, a designer should<br />

also be aware of the freedom in the<br />

arbitrary shaping of aluminium cross<br />

sections as enabled by the extrusion<br />

process (Fig. 1).<br />

This implies that with design and calculation<br />

in aluminium—more than<br />

with other materials such as steel,<br />

concrete, wood—it is necessary to<br />

“think” in aluminium. In other words,<br />

the possibilities with aluminium,<br />

in particular the freedom of shaping<br />

by extrusion, must be kept in<br />

mind from the very beginning of<br />

a design. 1–8<br />

Overview of Research<br />

Activities<br />

State of the Art<br />

In the past four decades, significant<br />

research has been carried out on aluminium<br />

technology as well as on its<br />

structural behaviour in Europe and<br />

elsewhere, such as in the United States.<br />

This has resulted in the development<br />

of new alloys, improvement of material<br />

properties, combination of materials,<br />

new joining techniques, improved<br />

strength, stability and fire resistance.<br />

Many joint industry projects have<br />

been undertaken and many research<br />

committees have been involved on<br />

national and international basis. In<br />

Europe, Technical Committee 2 of the<br />

ECCS, the European Convention for<br />

Constructional Steelwork, has been<br />

particularly active. 9<br />

In the early 1970s, research priorities<br />

were set, and joint industry research<br />

work commenced. Some of the first<br />

topics tackled were strength, plasticity<br />

and ductility in order to study<br />

the application of plastic analysis<br />

for aluminium alloys. 10–13 Top priority<br />

was given to stability, more precisely<br />

global and local buckling, and<br />

this subject still requires substantial<br />

attention. 14–26<br />

Another important topic in Europe<br />

and elsewhere was connections, where<br />

the structural behaviour of both preloaded<br />

and non-preloaded bolts was<br />

investigated. 27–31 In The Netherlands<br />

in particular and also elsewhere, the<br />

structural behaviour of welded connections<br />

was studied. 32–36 Subsequently,<br />

adhesive-bonded connections were<br />

investigated to look after their applicability<br />

in building and civil engineering.<br />

37,38 High priority was also given<br />

to the topic of fatigue because of the<br />

higher vulnerability of aluminium<br />

structures to fatigue as compared<br />

to steel structures. 39–51 This work is<br />

still in progress in many countries;<br />

in Europe it has been coordinated<br />

within the ECCS-TC2. As in the case<br />

of steel, for aluminium, most attention<br />

was given to the fatigue behaviour of<br />

welded details (Fig. 2). Research was<br />

also done on aluminium shell structures,<br />

which resulted in design rules<br />

as given in Eurocode 9, Part 1–5, Shell<br />

Structures. 52–54<br />

Forthcoming Activities<br />

The above-mentioned knowledge of<br />

structural behaviour must be completed<br />

and extended. New designs and<br />

applications demand improved design<br />

rules, which in turn demands further<br />

study of specific topics.<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: Sept. 15, 2009<br />

Paper accepted: July 22, <strong>2010</strong><br />

Fig. 1: Aluminium extruded sections<br />

Fig. 2: Fatigue loading of welded<br />

aluminium bridge<br />

430 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Some topics—related to the structural<br />

behaviour of aluminium—that require<br />

further study are as follows:<br />

– Research to optimise the loadbearing<br />

resistance of aluminium<br />

cross-sections, 55 which will further<br />

improve the existing design rules in<br />

Eurocode 9. Important aspects in<br />

this research are local and distortional<br />

buckling of thin-walled cross<br />

sections.<br />

– Research on “new” joining methods,<br />

such as “Friction Stir” welding 56–59<br />

and adhesive bonding. 60 For the<br />

latter, in particular, reliability and<br />

long-term behaviour (durability)<br />

are important aspects. This research<br />

will facilitate design rules for these<br />

joining technologies in Eurocode<br />

9, as well as in other national and<br />

international standards.<br />

– Research on improved methods to<br />

accurately determine the fatigue<br />

behaviour of aluminium components;<br />

except for existing design<br />

methods like safe life design and<br />

damage-tolerant design, the latter<br />

based on fracture mechanics, a promising<br />

continuum damage mechanics<br />

approach is under development.<br />

– Research on the structural behaviour<br />

of aluminium at elevated<br />

temperatures such as fire conditions.<br />

61–67 Although the strength<br />

and stiffness of aluminium decrease<br />

very fast at elevated temperatures,<br />

some aspects appear to be less sensitive<br />

at elevated temperatures compared<br />

to room temperature, such as<br />

buckling.<br />

Applied research such as that indicated<br />

above will result in better design rules,<br />

as has been shown with the research<br />

carried out in the past decades.<br />

In addition to extending the knowledge<br />

on the structural behaviour of aluminium,<br />

the designer must be instructed<br />

on how to apply this knowledge, that<br />

is, a type of knowledge transfer which<br />

must occur. This can be done through<br />

design projects where research institutions<br />

and industrial partners cooperate<br />

in order to demonstrate how to arrive<br />

at an optimum design in aluminium.<br />

This is particularly important for aluminium<br />

because aluminium is a rather<br />

young structural material as compared<br />

to steel, concrete, or bricks.<br />

Development of Design<br />

Standards<br />

The significant amount of research<br />

performed on the structural behaviour<br />

of aluminium thus far has culminated<br />

in design rules for structural applications.<br />

Many countries in Europe have<br />

an aluminium standard for building<br />

and civil engineering applications.<br />

However, most of them are quite outdated.<br />

The most recently updated ones<br />

are the British Standard BS 8118 68 and<br />

the Dutch Standard NEN 6710, 69 but<br />

these standards date from 1991; the<br />

old German Standard DIN 4113 70 was<br />

updated in 2002.<br />

As mentioned above, significant<br />

research in Europe was carried out<br />

under the umbrella of ECCS, the<br />

European Convention for Constructional<br />

Steelwork, where the<br />

Technical Committee 2 dealt with<br />

aluminium research that led in 1978<br />

to the first edition of European<br />

Recommendations. 9 Within the ECCS-<br />

TC2, research has continued since then,<br />

marking the beginning of a European<br />

Standard for the design of aluminium<br />

structures, Eurocode 9. 71 Although<br />

design rules in other national standards<br />

have been updated, Eurocode<br />

9 is by far the most comprehensive<br />

and up to date structural aluminium<br />

standard. 72–75<br />

The standard is subdivided into five<br />

parts: Part 1–1: General Structural<br />

Rules, Part 1–2: Fire Design, Part 1–3:<br />

Fatigue Design, Part 1–4: Cold-Formed<br />

Sheets, and Part 1–5: Shell Structures.<br />

Also, outside of Europe several design<br />

standards have been established in<br />

the past decades, for example, in the<br />

United States, Japan and Australia.<br />

Structural Design Applications<br />

The most important structural applications<br />

of aluminium can be found<br />

in transport and in building as well<br />

as civil engineering. In this paper,<br />

only building and civil engineering<br />

applications will be dealt with. The<br />

applications can be distinguished<br />

between offshore and onshore structures.<br />

Offshore applications are, for<br />

instance, helidecks, living modules,<br />

gangways, stairs, etc. Onshore applications<br />

are, for example, long-span<br />

roofs (space frames and domes),<br />

bridges, bridge decks, traffic gantries<br />

and sewage plants. A number<br />

of recently built structural applications<br />

of aluminium are reviewed by<br />

Soetens, 76 while many more and also<br />

older structural applications are dealt<br />

with by Mazzolani. 77 Some examples<br />

of recently built aluminium structures<br />

in The Netherlands are dealt with in<br />

the following sections.<br />

Aluminium Office Building<br />

An all-aluminium office building was<br />

built in Houten, The Netherlands,<br />

in 2000. It is the new office of the<br />

Aluminium Centre, the branch<br />

organisation of the Dutch aluminium<br />

industry. 78<br />

The concept of an all-aluminium<br />

office building, supported by hundreds<br />

of slender aluminium columns and<br />

designed by a Dutch architect, was the<br />

winning design in a contest and was<br />

chosen out of 64 designs. The architect<br />

was inspired by the rural landscape<br />

and called his design “aluminium forest”.<br />

The design is exceptional in that<br />

it is a one-storey building of about<br />

1000 m 2 supported by 380 aluminium<br />

columns (Fig. 3). Moreover, the stability<br />

of the system had to be provided<br />

by the columns as such; no additional<br />

bracings were allowed by the architect.<br />

The columns had a length of 6 m and<br />

diameters varying from 90 to 210 mm.<br />

Numerical simulations were carried<br />

out to investigate strength, stability<br />

and deformations of various models of<br />

the structure to decide upon the final<br />

design of the load-bearing system.<br />

To investigate the structural behaviour<br />

of the building system, a threedimensional<br />

finite element model of<br />

the entire system was developed. In<br />

all, six different designs were investigated.<br />

With the final design, sufficient<br />

stiffness (reduced deformations combined<br />

with an eigenfrequency above<br />

1 Hz) as well as sufficient stability and<br />

strength were achieved, among others,<br />

by introducing a small inclination to 80<br />

columns with a diameter of 210 mm. 78<br />

Aluminium Bridges<br />

In The Netherlands, the first aluminium<br />

bridge was installed in Amsterdam<br />

in 1955, followed by a few more aluminium<br />

bridges and/or bridge decks<br />

around the same time period, in different<br />

locations. In addition to the<br />

favourable properties of aluminium,<br />

as mentioned in the “Introduction”<br />

section, the supply of recycled aluminium<br />

after the 1945 war was one of the<br />

main reasons for these applications. In<br />

Europe and in North America only a<br />

small number of aluminium bridges<br />

have been built between the 1950s<br />

and the 1990s, probably because of the<br />

limited knowledge of the structural<br />

behaviour of aluminium.<br />

However, recently new initiatives have<br />

been taken and a “renewed” interest<br />

in aluminium bridges has arisen. 79,80<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 431


Fig. 5: Residential area traffic bridge<br />

Fig. 3: Aluminium office building<br />

In The Netherlands, four application<br />

areas of interest were selected:<br />

– movable bridges; 81<br />

– residential area bridges; traffic but<br />

also pedestrian bridges; 82<br />

– extension of existing bridges; 83<br />

– renovation of bridge decks. 84<br />

Attention was given to all four areas,<br />

but most attention was paid to the movable<br />

bridge. A specific type of movable<br />

bridge is shown here: a non-balanced,<br />

hydraulically driven traffic bridge, with<br />

18 m span and 12 m width.<br />

A key requirement for the above type<br />

of movable bridge is light weight. In<br />

the meantime, many bridge projects<br />

have been designed and built in all<br />

four application areas, that is, one<br />

movable traffic bridge in Amsterdam,<br />

heavily loaded by trucks which necessitated<br />

a thorough fatigue analysis<br />

(Fig. 4); several residential area traffic<br />

bridges (Fig. 5); many pedestrian<br />

bridges with architecturally pleasing<br />

designs (Fig. 6); an aluminium extension<br />

of a steel bridge whereby the<br />

2,5 m-wide concrete pedestrian lane<br />

was replaced by 4,8 m wide prefabricated<br />

aluminium sections on the<br />

existing steel consoles (Fig. 7); and<br />

finally a number of aluminium decks<br />

replacing bridge decks (steel, concrete,<br />

wood). Figure 8 shows 600 m 2 of the<br />

wooden bridge deck of the movable<br />

bridge in the A29 highway replaced<br />

by prefabricated aluminium panels<br />

mounted on the existing steel beams.<br />

This system was also employed in the<br />

Fig. 4: Hydraulically driven movable<br />

bridge in Amsterdam<br />

renovation of two steel bridges in<br />

Kentucky, USA.<br />

In order to provide an alternative route<br />

in the case of road works on a bridge<br />

or on a road along a canal, a floating<br />

roadway would be particularly useful.<br />

Therefore a single-lane, floating roadway<br />

for cars travelling up to 80 km/h<br />

has been designed, 85 built and tested<br />

to full scale (Fig. 9). The floating road<br />

consists of aluminium pontoons, built<br />

out of a box frame of welded extruded<br />

sections, aluminium side walls and bottom<br />

plates. The pontoons had typical<br />

dimensions of 5,3 × 3,5 × 1,5 m (length,<br />

Fig. 6: Movable pedestrian bridge near<br />

central railway station Amsterdam<br />

Fig. 7: Aluminium extension of a steel<br />

bridge<br />

width, height). These dimensions and<br />

its light weight (about 2500 kg) allowed<br />

easy transportation by trucks as well as<br />

easy assembly and disassembly.<br />

In addition to the low weight, corrosion<br />

resistance, whereby surface protection<br />

was not necessary, and low maintenance<br />

costs were important criteria for<br />

the choice of aluminium in the above<br />

bridge applications.<br />

An interesting 50-year-old aluminium<br />

traffic bridge in the Ruhr area of<br />

Germany deserves to be mentioned.<br />

Owing to the heavy industrial atmosphere<br />

in the Ruhr area, it is inspected<br />

regularly whereby the latest tests and<br />

inspections carried out in 2003 showed<br />

that the bridge is still in a good condition.<br />

86 Furthermore, in Northern<br />

Europe, in particular in Norway and<br />

Sweden, a number of aluminium<br />

bridges and many bridge decks have<br />

432 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Fig. 8: Wooden bridge deck replaced by aluminium sections<br />

Fig. 9: Aluminium floating road, full-scale testing<br />

been built in the last two decades. 87<br />

Apart from Europe, the concept of<br />

replacing existing bridge decks with<br />

aluminium decks is also considered in<br />

the United States. 88 Other aluminium<br />

structural applications are also receiving<br />

renewed attention. 89,90<br />

Concluding Remarks and<br />

Outlook<br />

Research carried out in the area of aluminium<br />

structures in the past decades is<br />

now starting to pay off. The knowledge<br />

on the structural behaviour as well as<br />

the up to date design rules allows better<br />

design of aluminium structures, in<br />

addition to avoiding stability problems<br />

and, in case of cyclically loaded structures,<br />

fatigue problems. Aluminium is<br />

a relatively young construction material<br />

as compared with concrete, steel,<br />

wood and masonry. The number of aluminium<br />

applications in load- bearing<br />

structures is still limited but, as has<br />

been reviewed above, there is a growing<br />

interest in the use of aluminium in such<br />

structures.<br />

Owing to the tendency to use lightweight<br />

and sustainable structures, the<br />

outlook for aluminium applications<br />

seems very positive. Enormous quantities<br />

of aluminium are available in<br />

the earth’s crust, and the opportunities<br />

for recycling are also very high,<br />

so there appears to be no limit to its<br />

application. This, combined with the<br />

availability of up to date design rules,<br />

promises a bright future for aluminium<br />

structural applications, provided the<br />

designer takes advantage of the favourable<br />

properties of aluminium and finds<br />

proper solutions to overcome the less<br />

favourable properties.<br />

Apart from the future probable<br />

research activities that have been<br />

described earlier in the “Overview of<br />

Research Activities” section, further<br />

areas of development are as follows:<br />

– material technology (new alloys,<br />

fibre metal laminates, self-healing<br />

materials, etc.);<br />

– structural applications of aluminium<br />

combined with other structural<br />

materials;<br />

– architectural design for building and<br />

civil engineering applications, which<br />

will take on more significance.<br />

This also means that the state of the<br />

art knowledge must be transferred to<br />

practice, to designers active in the field.<br />

One way to do this is to carry out joint<br />

industry projects where both research<br />

institutions and industrial partners<br />

take part. The European Aluminium<br />

Association (EAA) has taken an<br />

important initiative to contribute to<br />

this knowledge transfer: an e-learning<br />

tool—AluMATTER, in particular the<br />

“Structural applications module”—has<br />

been developed with interactive design<br />

examples based on Eurocode 9, which<br />

is of importance not only to practice<br />

but also for educational purposes.<br />

References<br />

[1] Mazzolani FM. Aluminium Alloy Structures,<br />

(2nd edn) E & FN SPON, Chapman & Hall:<br />

London, 1994.<br />

[2] Bulson PS (ed.), Aluminium Structural<br />

Analysis, Recent European Advances. Elsevier<br />

Applied Science Publishers, Barking, Essex,<br />

Great Britain, 1992.<br />

[3] Mazzolani FM (ed.), Aluminium Structural<br />

Design. CISM, Springer-Verlag: Wien, New York,<br />

2003.<br />

[4] Sharp ML. Behaviour and Design of<br />

Aluminium Structures. McGraw-Hill, Inc.:<br />

New York, 1992.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 433


[5] Kissell JR, Ferry RL. Aluminium Structures,<br />

A Guide to Their Specifications and Design. John<br />

Wiley & Sons, Inc.: New York, 1995.<br />

[6] Beaulieu D. Design of Aluminium Structures.<br />

Les presses de l’aluminium: Chicoutimi, Canada,<br />

2006.<br />

[7] Mazzolani FM. Design principles for aluminium<br />

structures, STAHLBAU SPEZIAL:<br />

Aluminium in Practice. Ernst & Sohn, Berlin,<br />

1998.<br />

[8] Mazzolani FM. Structural use of aluminium<br />

alloys in civil engineering. Proc. 2nd International<br />

Conference on Structural Engineering, Mechanics<br />

and Computation, Cape Town, South Africa,<br />

2004.<br />

[9] ECCS. European Recommendations for<br />

Aluminium Alloy Structures, 1st edn. Brussels,<br />

1978.<br />

[10] Mandara A, Mazzolani FM. Behavioural<br />

aspects and ductility demand of aluminium alloy<br />

structures. Proc. ICSAS Conference, Istanbul,<br />

1995.<br />

[11] Mazzolani FM, Piluso V. Prediction of rotation<br />

capacity of aluminium alloy beams. Proc.<br />

ICSAS Conference, Istanbul, 1995.<br />

[12] de Matteis G, Moen LA, Langseth M,<br />

Landolfo R, Hopperstad OS, Mazzolani, FM.<br />

Cross-sectional classification for aluminium<br />

beams—parametric study. ASCE, J. Struct. Eng.<br />

2001; 127(3): 271–279.<br />

[13] Lademo O-G, Hopperstad OS, Langseth<br />

M. Evaluation of yield criteria and flow rules<br />

for aluminium alloys. Int. J. Plast. 1999; 15(2):<br />

191–208.<br />

[14] Langseth M, Hopperstad OS. Local buckling<br />

of square thin-walled aluminium structures.<br />

Thin-Walled Struct. 1997; 27(1): 117–126.<br />

[15] Hopperstad OS, Langseth M, Hanssen L.<br />

Ultimate compressive strength of plate elements<br />

in aluminium: correlation of finite element analyses<br />

and tests. Thin-Walled Struct. 1997; 29(1–4):<br />

31–46.<br />

[16] Hoglund T. Shear buckling of steel and<br />

aluminium plate girders. Thin-Walled Struct.<br />

1997; 29: 13–30.<br />

[17] Mazzolani FM, Faella C, Piluso V, Rizzano<br />

G. Experimental analysis of aluminium alloy<br />

SHS-members subjected to local buckling under<br />

uniform compression, 5 th Int. Colloquium on<br />

Structural Stability, SSRC, Rio de Janeiro, 1996.<br />

[18] Moen LA, Hopperstad OS, Langseth M,<br />

Landolfo R, Dematteis G, Mazzolani FM.<br />

Rotational capacity of aluminium beams under<br />

moment gradient. Experiments and numerical<br />

simulations. J. Struct. Eng. 1999; 125(8): 921–929.<br />

[19] Mazzolani FM, Faella C, Piluso V, Rizzano<br />

G. Local buckling of aluminium channels under<br />

uniform compression: experimental analysis,<br />

Proc. ICSAS Conference, Espoo, Finland, 1999.<br />

[20] Soetens F, Mennink J, Snijder HH, van Hove<br />

BWEM. Experimental work and FEM calculations<br />

on the interaction of local buckling modes<br />

in aluminium sections. Proc. SIMS Conference,<br />

Lissabon, September 2000.<br />

[21] Chapuis J, Galambos TV. Restrained<br />

crooked aluminium columns. J. Struct. Div.<br />

ASCE 1982; 108(ST3): 511–524.<br />

[22] Sharp ML. Longitudinal stiffeners for compression<br />

members. J. Struct. Div. ASCE 1966;<br />

92(ST5): 187–211.<br />

[23] Sharp ML. Strength of beams or columns<br />

with buckled elements J. Struct. Div. ASCE 1970;<br />

96(ST5): 1011–1015.<br />

[24] Hopperstad OS, Langseth M, Tryland T.<br />

Ultimate strength of aluminium alloy outstands<br />

in compression: experiments and simplified analysis.<br />

Thin-Walled Struct. 1999; 34(4): 279–294.<br />

[25] Aalberg A, Langseth M. Larsen PK Stiffened<br />

aluminium panels subjected to axial compression.<br />

Thin-Walled Struct. 2001; 39(10): 861–885.<br />

[26] Mennink J, Soetens F, Snijder HH, van Hove<br />

BWEM. Buckling experiments and FE-analyses<br />

on complex aluminium extrusions. ASSCCA<br />

Conference, Sydney, June 2003.<br />

[27] Soetens F. Connections in aluminium alloy<br />

structures. Proc. ICSAS Conference, Cardiff,<br />

1987.<br />

[28] Ramirez J. Aluminium structural connections,<br />

conventional slip factors in friction grip<br />

joints. Proc. ICSAS Conference, Cardiff, 1987.<br />

[29] Menzemer CC. Failure of bolted connections<br />

in an aluminium alloy. J. Mater. Eng.<br />

Perform. ASM 1999; 8(2).<br />

[30] May JE, Menzemer CC. Strength of bolted<br />

aluminium alloy tension members. J. Struct. Eng.<br />

ASCE 2005; 131(7): 1125–1134.<br />

[31] Mazzolani FM, de Matteis G, Mandara A.<br />

Classification system for aluminium alloy connections,<br />

IABSE Colloquium, Istanbul, 1996.<br />

[32] Soetens F. Welded connections in aluminium<br />

alloy structures. HERON 1987; 32: 1–48.<br />

[33] Soetens F. Welded connections according<br />

to Eurocode 9 Design of Aluminium Structures.<br />

Proc. 7 th INALCO Conference, Cambridge<br />

(UK), April 1998.<br />

[34] Soetens F. Design of welded connections,<br />

STAHLBAU 67, Sonderheft Aluminium, Ernst<br />

& Sohn: Berlin, 1998.<br />

[35] Menzemer CC, Iasconne R. Reestablishing<br />

the shear strength of aluminium alloy fillet welds<br />

Weld. J. 2002; 81(4): 29–31.<br />

[36] Wang T, Hopperstad OS, Lademo O-G,<br />

Larsen PK. Finite element modelling of welded<br />

aluminium members subjected to four-point<br />

bending. Thin-Walled Struct. 2007; 45(3): 307–320.<br />

[37] Soetens F. Adhesive joints in aluminium alloy<br />

structures. Int. J. Adhes. Adhes. 1990; 10: 146–152.<br />

[38] Soetens, van Straalen IJJ, Dijkstra OD.<br />

Structural adhesive joints in aluminium. Third<br />

International Conference on Structural Adhesives<br />

in Engineering III, Bristol, June 1992.<br />

[39] Jaccard R, Kosteas D, Ondra R. Background<br />

Document to Fatigue Design Curves for Welded<br />

Aluminium Components, IIW doc. No. XIII-<br />

1588-95, 1995.<br />

[40] Kosteas D. On the fatigue behaviour of aluminium.<br />

STAHLBAU 67, Sonderheft Aluminium,<br />

Ernst and Sohn: Berlin, 1998.<br />

[41] Atzori B, Lazzarin P, Quaresimin M. A<br />

re-analysis on fatigue data of aluminium alloy<br />

bolted joints. Int. J. Fatigue 1997; 19(7): 579–588.<br />

[42] Sharp ML, Nordmark GE, Menzemer CC.<br />

Fatigue Design of Aluminium Components and<br />

Structures. McGraw-Hill, Inc.: New York, NY,<br />

1996.<br />

[43] Kosteas D, Jaccard R, Rothe J. Contribution<br />

to fatigue design rules in prEN 1999-1-3, R-ratio<br />

dependence, fatigue test factor and Simplified<br />

geometric stress concept. Proc. 9 th INALCO<br />

Conference, Cleveland, Ohio, USA, 2004.<br />

[44] Maddox SJ. Scale effect in fatigue of fillet<br />

welded aluminium alloys. Proc. 6 th INALCO<br />

Conference, Cleveland, Ohio, USA, 1995.<br />

[45] Soetens F. European research on fatigue<br />

of aluminium structures. Proc. 6 th INALCO<br />

Conference, Cleveland, Ohio, USA, 1995.<br />

[46] Maljaars J, Soetens F, van Straalen IJJ,<br />

van Hove BWEM. Fatigue tests on aluminium<br />

bridges. Proc. 9 th INALCO Conference,<br />

Cleveland, Ohio, USA, 2004.<br />

[47] Maljaars J, Soetens F, van Straalen, IJJ.<br />

Fatigue tests on aluminium bridges, HERON,<br />

Delft, 2005.<br />

[48] Maljaars J, Soetens F, Straalen IJJ. Fatigue<br />

of aluminium bridge decks, Struct. Eng. Int. 2006;<br />

16(4): 305–311.<br />

[49] Atzori B, Meneghetti G, Susmel L. A new<br />

procedure for the evaluation of fatigue strength<br />

of notched components applied to the evaluation<br />

of scale effect in aluminium alloy weldments.<br />

Proc. 8 th INALCO Conference, Munich, 2001.<br />

[50] Haagensen P, Oma S. Fatigue life improvement<br />

of welded aluminium alloys by peening.<br />

Proc. 8 th INALCO Conference, Munich, 2001.<br />

[51] Orjasaeter O, Skallerud B. Fatigue of wide<br />

plate weldments. Proc. 8 th INALCO Conference,<br />

Munich, 2001.<br />

[52] Mandara A, Mazzolani FM. On the stability<br />

of aluminium alloy cylindrical shells under<br />

axial compression. Costruzioni Metalliche 1989;<br />

2–3.<br />

[53] Mandara A, Mazzolani FM. Testing results<br />

and design procedures for axially loaded aluminium<br />

alloy cylinders. Proc. International<br />

Colloquium on Stability of Steel Structures,<br />

Budapest, 1990.<br />

[54] Mazzolani FM, Mandara A. Buckling of<br />

aluminium shells: proposal for European curves.<br />

Proc. 4 th International Conference on Thin-<br />

Walled Structures, Loughborough (UK), 2004.<br />

[55] Kutanova N, Soetens F, Snijder HH. Crosssectional<br />

instabilities of aluminium structural<br />

elements. Proc. CIMS Conference, Sydney, June<br />

2008.<br />

[56] Thomas WM. Friction stir welding and<br />

related friction process characteristics. Proc. 7 th<br />

INALCO Conference, Cambridge (UK), 1998.<br />

[57] Okura I, Naruo M, Vigh LG, Hagisawa N,<br />

Toda H. Fatigue of aluminium decks fabricated<br />

by friction stir welding. Proc. 8 th INALCO<br />

Conference, Munich, 2001.<br />

[58] Hori H, Tanikawa H, Seo N, Namba K.<br />

Tensile properties of friction stir welded joints—<br />

studies on characteristics of friction stir welded<br />

joints in structural aluminium alloys (report 3).<br />

Proc. 10 th INALCO Conference, Osaka, 2007.<br />

[59] Thomas WM, Verhaeghe G, Martin J. Staines<br />

DG, Stanhope C. Friction stir welding-process<br />

variants—an update. Proc. 10 th INALCO<br />

Conference, Osaka, 2007.<br />

434 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


[60] van Hove BWEM, Soetens F. Analytical<br />

and experimental research on strength properties<br />

of adhesive bonded joints in aluminium<br />

structures. Proc. 10 th INALCO Conference,<br />

Osaka, 2007.<br />

[61] Maljaars J, Fellinger JHH, Soetens F. Fire<br />

exposed aluminium structures, HERON 2005;<br />

50(4): 261–271.<br />

[62] Maljaars J, Soetens F, Katgerman L.<br />

Constitutive model for aluminium alloys<br />

exposed to fire conditions Metall. Mater. Trans.<br />

A 2008; 39: 778–789.<br />

[63] Langhelle NK, Eberg E, Amdahl J,<br />

Lundberg S. Buckling tests of aluminium columns<br />

at elevated temperatures. Proc. 15 th<br />

Offshore Mechanics and Arctic Engineering<br />

Conference, Firenze, Italy, 1996.<br />

[64] Maljaars J, Twilt L, Soetens F. Flexural buckling<br />

of fire exposed aluminium columns. Fire<br />

Safety J. 2009; 44: 711–717.<br />

[65] Maljaars J, Soetens F. Strength of MIG<br />

welded connections in fire exposed aluminium<br />

structures. J. Adv. Steel Constr. 2009; 5: 136–150.<br />

[66] Maljaars J, Soetens F, Snijder HH, Local<br />

buckling of aluminium structures exposed<br />

to fire. Part 1: Tests and Part 2: Finite element<br />

models. Thin-Walled Struct 2009; 47:<br />

1404–1428.<br />

[67] Maljaars J, Soetens F, Snijder HH. Local<br />

buckling of fire exposed aluminium members—<br />

a new design model. J. Struct. Eng., <strong>2010</strong>; 136:<br />

66–75.<br />

[68] BS 8118. Structural Use of Aluminium, Parts<br />

1 and 2. BSI: London, 1991.<br />

[69] NEN 6710. TGB 1990 Aluminium<br />

Constructies. NNI: Delft, 1991.<br />

[70] DIN 4113. Aluminiumkonstruktionen unter<br />

vorwiegend ruhender Belastung, Teil 1 und Teil 2.<br />

DIN: Berlin, 2002.<br />

[71] EN 1999. Eurocode 9: Design of aluminium<br />

structures, Parts 1-1 to 1-5. CEN: Brussels,<br />

2007.<br />

[72] Mazzolani FM, Mandara A, Langseth M.<br />

Plastic design of aluminium members according<br />

to EC9. Proc. International Conference on<br />

Steel and Aluminium Structures, Espoo, Finland,<br />

1999.<br />

[73] de Matteis G, Mandara A, Mazzolani FM.<br />

Calculation methods for aluminium T-stubs: a<br />

revision of EC9, Annex J. Proc. 8th INALCO<br />

Conference, Munich, 2001.<br />

[74] Mazzolani FM. Conception des structures<br />

en alliage d’ aluminium selon l’Eurocode 9(EN<br />

1999). Les Eurocodes—Conception des bâtiments<br />

et des ouvrages de genie civil, Editions Le<br />

Moniteur, Paris, 2005.<br />

[75] Soetens F. van Hove BWEM. Design<br />

of welded connections in aluminium structures.<br />

IABSE, Struct. Eng. Int. 2006; 16(4):<br />

326–331.<br />

[76] Soetens F, Structural applications of aluminium:<br />

an overview. Proc. 10 th INALCO<br />

Conference, Osaka, 2007.<br />

[77] Mazzolani FM. Structural applications of<br />

aluminium in civil engineering. IABSE, Struct.<br />

Eng. Int. 2006; 4: 280–286.<br />

[78] Soetens F, van Hove BWEM, Mennink J.<br />

Dutch aluminium technology office building<br />

in the Netherlands. Proc. Aluminium 2000<br />

Conference, Essen, Germany, 2000.<br />

[79] Soetens F, Maljaars J, de Kluyver D.<br />

Aluminium bridges and bridge decks—an<br />

overview of recent applications. Proc. 3rd<br />

International Conference in Building and<br />

Construction, Moscow, 2008.<br />

[80] Siwowski T. Aluminium bridges—past,<br />

present and future. Struct. Eng. Int. 2006; 16(4):<br />

319–326.<br />

[81] Soetens F, van Hove BWEM, Mennink J.<br />

Aluminium bridges in The Netherlands. Proc. 8 th<br />

INALCO Conference, Munich, 2001.<br />

[82] Soetens F, Mennink J, van Hove BWEM.<br />

Aluminium bridges, actual designs and prospects.<br />

Proc. ASSCCA Conference, Sydney, 2003.<br />

[83] Maljaars J, Soetens F, Burggraaf H, Design<br />

of an aluminium bicycle path integrated in a<br />

steel bridge. Proc. IABSE Symposium, Weimar,<br />

Germany, 2007.<br />

[84] Maljaars J, Soetens F, de Kluyver D. Structural<br />

design of aluminium bridge decks for existing<br />

bridges. Proc. IABSE Symposium, Chicago, 2008.<br />

[85] Soetens F, van Hove BWEM, Maljaars J,<br />

Janssen EGON, Mennink J. Floating aluminium<br />

roads. Proc. 9 th INALCO Conference, Cleveland,<br />

Ohio, USA, 2004.<br />

[86] Mader W, Pieper A. Schwansbell bridge<br />

celebrating 50 th birthday, Struct. Eng. Int. 2006;<br />

16(4): 356–360.<br />

[87] Hoglund T, Nilsson L. Aluminium in bridge<br />

decks and in a new military bridge in Sweden.<br />

Struct. Eng. Int. 2006; 16(4): 348–352.<br />

[88] Kissell JR, Das SK, Kaufman G.<br />

Opportunities for extruded aluminium alloy<br />

bridge decks for existing bridges. Light Metal<br />

Age 2008; 66(3): 12–18.<br />

[89] Kissell JR. Recent Developments in<br />

Aluminium Structural Design. Aluminium Now,<br />

Aluminium Association: Washington, DC, 2002.<br />

[90] Kissell JR, Skillingberg MH. Aluminium in<br />

construction applications. Aluminium Int. Today<br />

2006; 18(2): 25–26.<br />

Further Information<br />

1. http:// www.eaa.net/en/education/<br />

AluMATTER.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 435


Glass Tensegrity Trusses<br />

Maurizio Froli, Prof.; Leonardo Lani, Dr-Ing.; Department of Civil Engineering, University of Pisa,<br />

Pisa, Italy. Contact: l.lani@ing.unipi.it<br />

Abstract<br />

High transparency and modularity, retarded first cracking, non-brittle collapse<br />

and fail-safe design were the basic requirements that inspired and guided the<br />

development of a new kind of glass beams. The two basic conceptual design<br />

goals were to avoid any cracking at service and to get a ductile behaviour at<br />

failure. These objectives were reached by a preliminary subdivision of the beam<br />

into many small triangular laminated panes and by assembling them together by<br />

means of prestressed steel cables. Two prototypes have been constructed at the<br />

University of Pisa, tested in the elastic domain under dynamics loads and successively<br />

brought to collapse under quasi-static, increasing load cycles. In order to<br />

investigate the decay process of residual mechanical resources, the second prototype<br />

has been repaired twice by substituting just the damaged triangular panes<br />

and then tested again each time up to failure. Experimental results resulted in a<br />

good agreement with non-linear numerical simulations performed by appropriate<br />

finite element modelling.<br />

Keywords: structural glass; prestressed glass structures; post-breakage behaviour;<br />

structural ductility; fail-safe design; chemical tempering; fracture mechanics.<br />

Introduction<br />

Ductility is usually associated with<br />

metallic materials that are capable of<br />

developing large plastic deformations.<br />

On the other hand, fragility is traditionally<br />

associated with glass materials<br />

or ceramics.<br />

Nevertheless, some outstanding and<br />

innovative glass structures, like the<br />

Haus Pavilion in Rheinbach the<br />

Yurakucho canopy in Tokyo, the<br />

glass stairs of the Apple Stores in San<br />

Francisco, have been built even in seismic<br />

areas where fragile failures must<br />

be definitely avoided.<br />

Indeed, although glass is fragile and<br />

weakly tension resistant, it has a very<br />

high compressive strength and, if conveniently<br />

connected with other ductile<br />

materials, for example, by means<br />

of gluing or by prestressing, it is able<br />

to form composite structures of high<br />

mechanical performances and also<br />

global ductility.<br />

It is well known that stress concentrations<br />

that occur at the apex of<br />

microscopic surface cracks, always<br />

present even in virgin specimens, are<br />

responsible for the intrinsic fragility<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: March 04, <strong>2010</strong><br />

Paper accepted: June 29, <strong>2010</strong><br />

of glass 1 and for its relative low tensile<br />

strength. An apparent higher tensile<br />

strength is obtained by thermal or<br />

chemical tempering treatments, which<br />

induce surface compression stresses<br />

that inhibit crack initiation and propagation<br />

but do not exert any influence<br />

on fragility. 2<br />

Prestressed Composite<br />

Glass Beams<br />

Basic Concepts<br />

The intrinsic fragility of glass may be<br />

overcome by organizing the whole<br />

structure in two or more hierarchic<br />

levels, each of them composed of a<br />

parallel, redundant assemblage of at<br />

least two structural components.<br />

The hierarchic organization of the<br />

components ensures that the sequence<br />

of progressive damage follows a preestablished<br />

order starting from the<br />

level where the weakest components<br />

are. Therefore, if we put ductile materials<br />

at the lowest level, we can be<br />

sure that the failure process will start<br />

from here accompanied by large plastic<br />

deformations, that is, with a global<br />

ductile behaviour.<br />

On the other hand, redundancy<br />

ensures at each level that, when a single<br />

component fails, the other partner<br />

components are still able to bear the<br />

load although with a reduced degree<br />

of safety. 3 In laminated glass panes,<br />

the application of this principle also<br />

ensures a pseudo-ductile behaviour: it<br />

is known that if a glass sheet breaks,<br />

the other sheets are still able to bear<br />

the load, and even if all the sheets<br />

break into large fragments, (only fully<br />

thermally tempered glass breaks into<br />

many small fragments), the redundant<br />

sandwich structure assures a postbreakage<br />

stiffness and bearing capacity<br />

of the component that is similar, to<br />

some extent, to material ductility. 4<br />

For these reasons, a suitable application<br />

of the two basic principles of hierarchy<br />

and redundancy can provide a<br />

structure with decisive properties of<br />

global ductility and fail-safe design<br />

even if mostly composed of glass components.<br />

Fig. 1 shows the structural<br />

organization of the present type of<br />

glass beams.<br />

Additionally, if the integrity of the<br />

structure is enhanced by prestressing,<br />

compression stresses superimpose in<br />

glass elements to those produced by<br />

tempering, thus increasing the apparent<br />

tensile strength of the material.<br />

Structural Conceptual Design<br />

of Trabes Vitreae Tensegrity Beams<br />

Experiments reveal that when a traditional<br />

glass beam is submitted to<br />

increasing flexural loads, it cracks at<br />

a certain load, developing characteristic<br />

crack patterns. To avoid an uncontrolled<br />

process of crack initiation and<br />

propagation, the idea considered was<br />

to govern it by regularly pre-cutting<br />

the glass surface into many equilateral<br />

triangle panes and connecting them<br />

together using a system of prestressed<br />

steel cables.<br />

The principle of tensegrity permeates<br />

this concept, therefore it was decided<br />

to call these beams Trabes Vitreae<br />

Tensegrity or TVT, mixing Latin and<br />

English words.<br />

Each triangular pane is composed of<br />

two 5 mm thick chemically tempered<br />

glass sheets 5 laminated by means of<br />

a 1,52 mm thick PolyVinylButyral<br />

(PVB) interlayer.<br />

The beam is composed of two parallel<br />

twin curtains 174 mm apart, braced<br />

on the upper side by a horizontal truss<br />

436 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


Redudancy, parallel assemblies<br />

Curtain A Curtain B<br />

Steel<br />

cable<br />

Steel<br />

cable<br />

Fig. 1: Hierarchy and redundancy organization<br />

Phase 0: Pure Prestress<br />

The structural behaviour of TVT<br />

beams is analogous to that of segmenand<br />

connected together at the lower<br />

edge nodes by means of hollow stainless<br />

steel structures (Figs. 2 and 3).<br />

Each curtain is made of a Warrenlike<br />

appearance of glass panes (Fig. 2)<br />

jointed at the apex by means of stainless<br />

steel nodes (Fig. 3). Mechanical<br />

bolting between the glass panes and<br />

steel nodes was avoided as dangerous<br />

local tensile peaks always occur in<br />

glass holes.<br />

Instead, the nodes are mutually<br />

connected by means of AISI 304 stainless<br />

steel cables tensioned by screw<br />

tighteners. Consequently, only contact<br />

pressure is exchanged between glass<br />

and steel nodes due to the prestress<br />

action. In order to attenuate local<br />

contact peaks, the vertices of the glass<br />

panels are round, and 1 mm thick<br />

AW-1050A grade aluminium alloy<br />

Glass sheet<br />

Glass sheet<br />

Glass sheet<br />

Glass sheet<br />

Laminated<br />

pane<br />

Laminated<br />

pane<br />

1st Level 2nd Level 3rd Level<br />

Hierarchy, increasing mechanical strength<br />

sheets are interposed between steel<br />

and glass.<br />

The redundancy principle is applied at<br />

two different levels: the first is that of<br />

the doubly laminated panes and the<br />

second that of the parallel arrangement<br />

of the two twin curtains of glass<br />

panes and steel cables as sketched in<br />

the scheme of Fig. 1. The relatively<br />

large spacing of the two curtains gives<br />

the beam an appreciable torsional<br />

stiffness and good lateral torsional<br />

buckling stability.<br />

Qualitative Structural Behaviour<br />

tal prestressed concrete beams. During<br />

the shop assemblage of a beam, the<br />

two twin curtains are placed on a<br />

horizontal plane and prestressed. The<br />

dead load of the curtains is entirely<br />

sustained by the surface, thus only<br />

prestress forces act, inducing a quasiisotropic<br />

distribution of compression<br />

stresses in the glass panes (Fig. 4).<br />

Phase 1: Glass Decompression<br />

When in service, under the flexural<br />

action of dead loads and increasing<br />

external loads, tension stresses in the<br />

lower parts of the glass panels gradually<br />

diminish prestress compressions<br />

until a limit state of decompression is<br />

reached in the central part of the beam.<br />

When the external loads are further<br />

increased, the decompression propagates<br />

from middle span towards the<br />

supports. This stage has been denoted<br />

as Phase 1—glass decompression.<br />

Since the steel nodes exert unilateral<br />

restraint only at the point of contact,<br />

the decompressed vertices of the glass<br />

panels detach and simply move a small<br />

distance from their supports without<br />

developing tension stresses. The static<br />

scheme of the beam changes thus into<br />

that sketched in Fig. 5 where flexural<br />

and shear tension forces are sustained<br />

respectively by the lower steel bars<br />

and one order of the diagonal steel<br />

bars. Compression stresses flux within<br />

the glass panels following typical<br />

“boomerang-shaped” patterns visible<br />

Fig. 2: The prototype TVTb beam<br />

Fig. 4: Phase 0 calculated compression isolines<br />

Fig. 3: Steel node<br />

Fig. 5: Phase 2 calculated compression isolines<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 437


1000,00 N<br />

1000,00 N<br />

1000,00 N<br />

1000,00 N<br />

Y<br />

X<br />

Fig. 6: Global model of TVTb<br />

in the same graph. Only secondary<br />

tension stresses of lower intensity<br />

affects glass.<br />

Phase 2: Buckling of Upper Cables<br />

Compressed steel cables are gradually<br />

de-tensioned: when the prestress load<br />

is fully compensated, they buckle away.<br />

This limit state has been denoted as<br />

Phase 2—buckling phase.<br />

Phase 3: Collapse<br />

After Phase 2 has been reached, a further<br />

increase in load leads to an augmentation<br />

of stress compressions in<br />

the glass panes and tension stresses in<br />

the steel rods. The dimensioning of the<br />

different component parts of the beam<br />

can be performed so that the final<br />

Phase 3—collapse takes place due to<br />

the yielding of the steel cables and not<br />

because of glass rupture in compression,<br />

thus resulting in a ductile collapse<br />

accompanied by large displacements.<br />

Depending on the slenderness of the<br />

steel cables and on the prestress intensity,<br />

Phase 2 may even follow Phase 3,<br />

as indeed happened in the prototype<br />

beam, and illustrated in the following<br />

text.<br />

Numerical Modelling<br />

Four different finite element model<br />

(FEM) analyses have been performed<br />

to predict the various aspects of the<br />

structural behaviour of the beam and<br />

to better calibrate the design of the<br />

prototypes (Fig. 6):<br />

– 2D non-linear geometrical analysis<br />

to evaluate the effect of prestress on<br />

the flexural stiffness of the beam;<br />

– 2D local buckling analysis to evaluate<br />

instability effects of prestress for<br />

each single glass pane;<br />

– 3D local geometric non-linear analysis<br />

to evaluate the transversal stiffness<br />

of the different joints;<br />

– 3D geometric non-linear analysis<br />

to evaluate the torsional stiffness of<br />

the beam.<br />

The glass panels have been modelled<br />

by Shell elements (4-node, 24 degrees<br />

of freedom) while the steel cables<br />

have been reproduced by Bar elements<br />

(2-node, 6 degrees of freedom)<br />

that react only to tension stresses. The<br />

constitutive laws for the two materials<br />

have been deduced from European<br />

Code 6 , that is, glass has been schematized<br />

as a linear brittle elastic material<br />

and stainless steel as linear elastic–<br />

plastic material with a linear hardening<br />

branch.<br />

In order to model contacts between<br />

glass panels and steel nodes a set of<br />

point contact elements was introduced<br />

that were capable of reacting just<br />

to compression stresses. Aluminium<br />

sheets were not included in the model<br />

because of their relatively small thickness<br />

and, for practical reasons, owing to<br />

the coincidence of the Young’s moduli<br />

of aluminium and glass. The transversal<br />

stiffness of the joint was preliminarily<br />

investigated by a 3D local model with<br />

solid elements (8-node, 24 degree of<br />

freedom) (Fig. 7).<br />

Calculations have substantially confirmed<br />

the intuitive predictions synthetically<br />

described in the section<br />

Qualitative Structural Behaviour with<br />

the only exception that the buckling<br />

phase of upper steel cables (Phase 2)<br />

does not influence significantly the<br />

flexural response of the beam (Fig. 8).<br />

On the other hand, the decompression<br />

phase (Phase 1) and the yielding phase<br />

(Phase 3) of the lower steel cables can<br />

be clearly recognized.<br />

Figure 9 shows the load factor versus<br />

displacement of the middle span point<br />

for different prestress (from 2 to 12<br />

kN) N p load in the steel cables, assuming<br />

P = 1 kN, the force applied to each<br />

of the eight nodes of the beam, corresponds<br />

to a load factor one. The first<br />

stiffness reduction is associated with<br />

the decompression of the lower part<br />

of the beam (Phase 1). By increasing<br />

the prestress level, the intensity of the<br />

external load that induces the decompression<br />

phase increases. The second<br />

step of stiffness decay is associated<br />

with the yielding of the lower steel<br />

cable but, of course, the ultimate limit<br />

load results independent of N p .<br />

Figure 10 shows the influence of N p<br />

on the axial force in the lower cable of<br />

the beam, and how the ultimate load is<br />

independent on the value of prestress.<br />

The hardening properties of stainless<br />

steel allowed the analysis to progress<br />

beyond the yield initiation of the lower<br />

bars until the buckling of the upper<br />

bars and of the middle span glass<br />

panels.<br />

438 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


It can be observed that the theoretical<br />

mechanical response of the beam is<br />

substantially bilinear until yielding of<br />

the steel bars occurs, while the buckling<br />

of the upper cables does not seem<br />

to have a relevant influence on the<br />

overall residual stiffness.<br />

Furthermore, the static principle,<br />

which states that prestress controls<br />

only serviceability limit states but not<br />

the ultimate limit states, is confirmed.<br />

(a)<br />

Z<br />

Y<br />

X<br />

Experimental Tests<br />

Virgin Specimens<br />

Y<br />

After the construction and testing of<br />

a first prototype (TVTa), which suffered<br />

some assemblage problems concerning<br />

prestress operation, a second<br />

prototype beam (TVTb) was prepared<br />

having a length of 3300 mm and a<br />

height of 572 mm. This prototype has<br />

been submitted to dynamic and quasistatic<br />

cyclical laboratory tests to completely<br />

characterize the experimental<br />

structural behaviour of the specimen<br />

and to compare it with numerical<br />

predictions.<br />

(b)<br />

Fig. 7: Joint model (a) and (b)<br />

Plate stress 22 Mid plane (MPa)<br />

−4<br />

−8<br />

−13<br />

−17<br />

−21<br />

−26<br />

−30<br />

−34<br />

−35<br />

Fig. 8: Calculated principal compression stresses<br />

Z<br />

X<br />

Dynamic Test<br />

Dynamic tests have been performed<br />

by inducing sudden impulses both<br />

in the horizontal and in the vertical<br />

direction at middle span. To produce<br />

the impulses, a mass of 34 kg was<br />

slowly applied to the beam by means<br />

of a steel rod thus inducing a state of<br />

initial distortion. As the rod was suddenly<br />

cut, the beam was submitted to<br />

damped free oscillations. Vertical and<br />

horizontal acceleration histories were<br />

recorded at some representative points<br />

of the beam, which allowed to evaluate<br />

eigen vibration periods and to control<br />

the attitude of the structure to damp<br />

free oscillations.<br />

Load factor<br />

9<br />

8<br />

7<br />

6<br />

Np = 2 kN<br />

Np = 4 kN<br />

5<br />

4<br />

Np = 6 kN<br />

Np = 8 kN<br />

Np = 10 kN<br />

3<br />

Np = 12 kN<br />

2<br />

1<br />

0<br />

−10 0 10 20 30 40 50 60 70 80 90<br />

Displacement (mm)<br />

Fig. 9: Load factor versus vertical displacement<br />

The assessment of the structural frequencies<br />

was deduced from the trend<br />

of the accelerations, appraising the distance<br />

between two consecutive maxima.<br />

Table 1 compares the theoretical<br />

and the experimental results.<br />

Quasi-Static Cyclic Tests<br />

The static test of TVTb prototype<br />

has been performed in two different<br />

stages: during the first one, the specimen<br />

was submitted to a progressively and<br />

cyclically increasing loading condition.<br />

In the second stage, the load was<br />

increased monotonically until the<br />

collapse of the beam took place at a<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 439


Load factor<br />

9<br />

Np = 2 kN<br />

8<br />

Np = 4 kN<br />

7<br />

Np = 6 kN<br />

6<br />

Np = 8 kN<br />

Np = 10 kN<br />

5<br />

Np = 12 kN<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 2000 4000 6000 8000 10 000 12 000 14 000 16 000 18 000 20 000<br />

Axial force (N)<br />

Fig. 10: Load factor versus axial force<br />

T est no. Form Measured frequency (Hz) Calculated frequency (Hz)<br />

1 In plane 19,1 76,9<br />

2 In plane 16,9 76,9<br />

1 Out of plane 13,4 12,1<br />

2 Out of plane 15 12,1<br />

Table 1: Results of the dynamics tests<br />

total applied load of 61,84 kN (load<br />

factor = 7,73).<br />

Before the application of external<br />

loads, vertical and horizontal displacements<br />

induced by self weight and prestress<br />

were measured over a few days.<br />

Such investigations led to the conclusion<br />

that tension and rigidity reductions<br />

that could occur with time due to<br />

the relaxation of the cables or to other<br />

viscosity phenomena in the PVB interlayer<br />

could be substantially neglected.<br />

The program of cyclic loading has<br />

been performed to check precedent<br />

intuitions and theoretical analyses and<br />

to evaluate the different structural<br />

resources of the prototype, namely, the<br />

ability to sustain repeated occurrence<br />

of increasing loads without damage<br />

or significant performance decay, the<br />

response to the influence of unavoidable<br />

geometrical imperfections and the<br />

eventual capacity to dissipate energy<br />

without damage.<br />

The results of the cyclic test program<br />

are represented in Fig. 11 in terms of<br />

total applied load versus middle span<br />

vertical displacement. The experimental<br />

results are compared in the same<br />

graph with the analytical monotonic<br />

response of the 2D FEM model.<br />

The comparison allows the conclusion<br />

that numerical and experimental<br />

9 72 kN<br />

8<br />

7<br />

6<br />

Numerical model<br />

Final pushover test<br />

64 kN<br />

56 kN<br />

48 kN<br />

results are rather close to each other<br />

with some limited discrepancies:<br />

– The first knee related to Phase 1<br />

(glass decompression) is recognizable<br />

although less marked as in<br />

numerical analysis;<br />

– The actual stiffness of the beam<br />

before the first decompression<br />

knee is lower than that obtained by<br />

numerical analysis;<br />

– At each new load cycle, the level of<br />

the decompression load decreases<br />

although the residual deformation is<br />

very limited.<br />

Each cycle of Fig. 11 encloses a finite<br />

area showing that the beam is also<br />

surprisingly able to dissipate energy<br />

before any damage occurs in the<br />

component materials. This can be<br />

attributed to the friction that develops<br />

due to relative slip movements at<br />

the interface between the glass and<br />

steel nodes and perhaps also to viscoelastic<br />

slip movements in the PVB<br />

interlayer.<br />

Very small transversal displacements<br />

were measured (not shown here for<br />

the sake of brevity) at each load cycle<br />

evidencing how good the torsional<br />

stiffness of the beam is and how it<br />

remains constant throughout the progression<br />

of the load cycles.<br />

After the completion of the cyclical<br />

load program, the beam has been submitted<br />

to a monotonic increasing load<br />

up to collapse. In the graph of Fig. 11,<br />

the load versus displacement curve of<br />

this stage (blue line) is compared with<br />

the theoretical curve. The experimental<br />

curve has almost no decompression<br />

knee but now the second knee is visible,<br />

corresponding to yielding of the<br />

lower steel bar, which occurred at a<br />

higher load level than predicted. First<br />

failure symptoms therefore manifested<br />

in the ductile component material of<br />

the composite structure (Fig. 12).<br />

Load factor<br />

5<br />

4<br />

3<br />

Cyclic tests<br />

40 kN<br />

32 kN<br />

24 kN<br />

2<br />

16 kN<br />

1<br />

8 kN<br />

0 0 kN<br />

0 10 20 30 40 50 60 70<br />

Displacement (mm)<br />

Fig. 11: Load of middle point versus vertical displacement compared with FEM<br />

(dotted line)<br />

Fig. 12: TVTβ at failure<br />

440 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


7<br />

56 kN<br />

Conclusion<br />

Load factor<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−5 0 5 10 15 20<br />

TVTβ<br />

Fig. 13: Displacement origins have been shifted to the right in TVTb bis/tris<br />

Owing to the hardening properties<br />

of stainless steel, the load can be<br />

increased even beyond yielding: the<br />

final collapse of the specimen was<br />

reached when the upper parts of middle<br />

span glass panels buckled away.<br />

Repaired Specimens<br />

After prototype TVTb completely<br />

collapsed as a consequence of the<br />

breaking of middle span panels, it was<br />

repaired by substituting just the broken<br />

panels. Prestress was restored at<br />

the same levels of the virgin specimen<br />

TVTb.<br />

The first repaired prototype was<br />

labeled as TVTbbis and submitted<br />

to the same increasing load cycles of<br />

TVTb.<br />

Since the clam plates of the central<br />

panels in TVTb resulted in a little dam-<br />

TVTβbis<br />

TVTβtris<br />

48 kN<br />

40 kN<br />

32 kN<br />

24 kN<br />

16 kN<br />

8 kN<br />

0 kN<br />

25 30 35 40 45 50 55 60 65<br />

Displacement (mm)<br />

age after buckling, they could not offer<br />

the same degree of restraint as the virgin<br />

ones.<br />

Therefore, the central panels of<br />

TVTb bis buckled corresponding to<br />

a load factor of 4 instead of the<br />

previous 7,5.<br />

Next, in prototype TVTb bis, the collapsed<br />

central panels were substituted.<br />

The second repaired prototype was<br />

called TVTbtris and exhibited almost<br />

the same ultimate load factor as the<br />

precedent version.<br />

Figure 13 shows the load factor versus<br />

middle span vertical displacement<br />

cycles of the virgin specimen and of<br />

the two repaired versions. A rather<br />

good retention of the stiffness properties<br />

of the repaired specimens can be<br />

observed together with a progressive<br />

increase in the dissipated energy.<br />

The numerical results and the test<br />

experiments on virgin TVTb prototype<br />

of a prestressed composite glass–steel<br />

beam allow the conclusion that the<br />

constructional principle is valid and<br />

merits further technological improvement<br />

and research work.<br />

The experimental and numerical<br />

results have underlined that TVT composite<br />

glass–steel beams are able to<br />

develop a ductile break-up and that the<br />

serviceability limit state is governed by<br />

the level of prestress in the steel cables.<br />

The cyclic load programme also evidenced<br />

that these glass beams are able<br />

to dissipate energy through friction<br />

and viscoelasticity without damage.<br />

The segmental, modular features and<br />

the tensile integrity of these beams<br />

allow substitutions to be limited just to<br />

the collapsed or cracked panels, thus<br />

reducing repair costs.<br />

References<br />

[1] Menčik J. Strength and Fracture of Glass and<br />

Ceramics. Elsevier: London, 1992.<br />

[2] Sedlacek G. Ein Bemessungskonzept zur<br />

Festigkeit thermisch vorgespannter Gläser.<br />

Shaker Verlag: Aachen, 2000.<br />

[3] Rice P, Dutton H. Structural Glass, 2nd edn.<br />

Spon Press: London, 2004.<br />

[4] Kott A, Vogel T. Safety of laminated glass<br />

structures after initial failure. IABSE Struct.<br />

Eng. Int. 2004; 14(2): 134–138).<br />

[5] Macrelli G. Process control methods for chemical<br />

strengthening of glass on industrial scale.<br />

Proc. XIX Int. Cong. Glass, Edinburgh, 2001.<br />

[6] EN572 – Glass in Building, Basic Soda Lime<br />

Silicate Glass Products, Definitions and General<br />

Physical and Mechanical Properties, CEN/TC<br />

129 Glass in Building, 2004.<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 441


A Simplified Serviceability Assessment of Footbridge<br />

Dynamic Behaviour Under Lateral Crowd Loading<br />

Luca Bruno, Associate Prof. and Fiammetta Venuti, Postdoc Fellow; Dept. of Structural and Geotechnical Engineering,<br />

Politecnico di Torino, Torino, Italy. Contact: fiammetta.venuti@polito.it<br />

Abstract<br />

In this paper a simplified approach to footbridge serviceability assessment under<br />

lateral crowd loading is proposed. The approach is based on the determination of<br />

a limit crowd density, which causes the lateral acceleration of the deck to reach<br />

the human perceptibility threshold. The approach relies on a lateral load model<br />

that describes the action of both uncorrelated and synchronised among-eachother<br />

pedestrians by proposing a law that links the percentage of synchronised<br />

pedestrians to the value of the crowd density, which is assumed to be uniformly<br />

distributed along the span. The approach is applied to three case studies and the<br />

results are compared with those obtained through the application of other serviceability<br />

and stability criteria.<br />

Keywords: serviceability criteria; footbridges; pedestrian load; synchronisation;<br />

lock-in.<br />

Introduction<br />

Over the last decade, the problem of<br />

footbridge serviceability under humaninduced<br />

excitation has been one of the<br />

important areas of research in the field<br />

of civil engineering. The interest of<br />

researchers and engineers is due to the<br />

recent trend towards the construction of<br />

more and more slender and lightweight<br />

footbridges, which are extremely prone<br />

to vibration. This interest is testified by<br />

an intense research activity (reviewed,<br />

e.g. in Refs. [1, 2]) and the publication of<br />

guidelines for the design of footbridges,<br />

especially focused on their dynamic<br />

behaviour under pedestrian loading. 3–6<br />

In particular, considerable attention<br />

has been directed towards the phenomenon<br />

of the so-called synchronous lateral<br />

excitation (SLE), first observed on<br />

the T-bridge in Japan 7 and brought to<br />

the world’s attention after the closure<br />

of the London Millennium Bridge in<br />

2000. 8 In recent years, SLE has been<br />

recognised to be affected by two<br />

kinds of synchronisation phenomena<br />

(e.g. Refs. [9, 10]): the pedestrian–<br />

structure (ps) one, also called lock-in,<br />

takes place when a pedestrian perceives<br />

the lateral motion of the walking surface<br />

and tends to synchronise his/her pacing<br />

Peer-reviewed by international experts<br />

and accepted for publication<br />

by <strong>SEI</strong> Editorial Board<br />

Paper received: December 01, 2009<br />

Paper accepted: May 28, <strong>2010</strong><br />

rate to that of the surface in order to<br />

maintain body balance; the pedestrian–<br />

pedestrian (pp) synchronisation arises<br />

when the pedestrian’s walking is constrained<br />

because of high crowd density<br />

values so that people tend to walk with<br />

the same frequency and a null relative<br />

phase angle, that is, they synchronise<br />

with one another. The complex relation<br />

between ps and pp synchronisation<br />

remains a subject of debate in the scientific<br />

community, especially in relation to<br />

their mutual role in the SLE triggering<br />

process. 11 Nevertheless, what is widely<br />

accepted is that the ps synchronisation<br />

cannot occur if the deck vibrations<br />

are below the pedestrian perceptibility<br />

threshold, while the pp synchronisation<br />

cannot take place if the crowd<br />

density is below the threshold for<br />

unconstrained free walking.<br />

Several criteria have been suggested to<br />

avoid occurrence of lock-in and guarantee<br />

pedestrian comfort. According<br />

to the authors, these can be roughly<br />

classified into two approaches corresponding<br />

to different stages of the SLE<br />

process in which the criterion is derived.<br />

According to the first approach, lock-in<br />

can be avoided if the amplitude of the<br />

lateral acceleration of the deck does<br />

not exceed a limit value corresponding<br />

to the pedestrian perceptibility threshold—for<br />

instance, the Sétra/AFGC 4<br />

and Synpex 5 technical guides recommend<br />

a limit value ̇ż lim =0,1m/s . The<br />

2<br />

second approach (e.g. Refs. [8, 12, 13])<br />

is derived once the lock-in has occurred<br />

(i.e. the lateral vibration is higher than<br />

the perceptibility threshold) by identifying<br />

two degrees of severity in the ps<br />

synchronisation process: a stable condition,<br />

where the lock-in has been triggered<br />

but the lateral response remains<br />

below a value of about 10 to 15 mm 13<br />

and an unstable condition, where the<br />

pedestrians exert larger lateral forces<br />

leading the structural response to an<br />

unstable amplification. The approach<br />

defines a limit number of pedestrians<br />

N lim as the one that sets off the unstable<br />

condition. For this reason, the criteria<br />

based on the second approach<br />

should be stability rather than serviceability<br />

criteria.<br />

Both approaches imply the need<br />

to define suitable load models that<br />

describe the load condition in the considered<br />

stage of the SLE process. The<br />

models related to the first approach<br />

consider the pre-lock-in stage and<br />

describe the pedestrian behaviour as<br />

random, that is, the pedestrians are<br />

assumed to walk with random phases<br />

(e.g. in Ref. [4]). On the contrary, the<br />

models related to the second approach<br />

rely on the assumption that lock-in has<br />

already occurred, that is, the pedestrian<br />

motion and force are synchronised<br />

and proportional to the lateral<br />

motion of the deck.<br />

According to the authors, the following<br />

main considerations can be<br />

outlined:<br />

– the pp synchronisation is not explicitly<br />

accounted for in none of the<br />

models, even though it could occur<br />

both in the pre- and post-lock-in<br />

stages, since it depends on the crowd<br />

density value;<br />

– it is not possible to establish which<br />

approach is the most conservative in<br />

absolute terms. In fact, the first one<br />

is the most conservative in terms of<br />

deck acceleration amplitude (the<br />

perceptibility threshold is certainly<br />

lower than the stability one) but this<br />

feature does not necessarily imply<br />

the same in terms of corresponding<br />

pedestrian number, because the<br />

lateral force of a single pedestrian is<br />

higher within the lock-in stage than<br />

in the pre-lock-in one.<br />

442 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


In this paper an alternative approach<br />

to footbridge serviceability assessment<br />

under lateral crowd loading is<br />

proposed. This approach is based on<br />

the consideration that the two concepts<br />

of limit values could be linked,<br />

that is, the limit number of pedestrians<br />

may be interpreted as the number<br />

of pedestrians that induce the<br />

limit value of the acceleration. The<br />

positive aspects of the two approaches<br />

are therefore retained: on one hand,<br />

the definition of a critical number of<br />

pedestrians could be useful for the<br />

designer and the footbridge owner in<br />

view of the development of control<br />

strategies to limit the flow of pedestrians;<br />

on the other hand, the analysis<br />

of the pre-lock-in stage is expected to<br />

be more relevant in order to predict<br />

and prevent the occurrence of SLE.<br />

The approach is accompanied by the<br />

proposal of a load model within the<br />

pre-lock-in stage that also accounts for<br />

the pp synchronisation due to crowd<br />

density. In the following, the proposed<br />

force model and comfort criterion are<br />

described and compared with other<br />

serviceability/stability criteria found<br />

in the literature through their application<br />

to some case studies.<br />

Proposed Approach<br />

The proposed approach is based on<br />

the derivation of a critical crowd density<br />

r lim , which causes the lateral acceleration<br />

of the deck to reach the limit<br />

value ̇ż lim =0,1m/s 2 . This limit value<br />

corresponds to the pedestrian perceptibility<br />

threshold recommended by<br />

Sétra/AFGC 4 and Butz et al. 5 , that is,<br />

to the upper bound of the pre-lock-in<br />

stage. The following assumptions have<br />

been adopted:<br />

1. The structural response can be estimated<br />

with sufficient accuracy using<br />

a single-degree-of-freedom (SDOF)<br />

modal equation for the mode of<br />

interest.<br />

2. The deck mass is constant along the<br />

footbridge span.<br />

3. The crowd is uniformly distributed<br />

along the span, so that both the<br />

crowd mass and the force exerted<br />

are constant in space. The force is<br />

applied on the deck so that its sign<br />

matches the sign of the deformed<br />

shape.<br />

4. The pre-lock-in stage is considered,<br />

that is, the pedestrians are not<br />

synchronised to the bridge motion,<br />

but they could be synchronised<br />

among each other owing to crowd<br />

density.<br />

It follows that the critical value of the<br />

crowd density r lim refers to pedestrians<br />

uniformly distributed along the<br />

footbridge walking path, that is, to a<br />

load condition widely adopted in the<br />

footbridge design and corresponding<br />

to a uniform crowd flow.<br />

The structural dynamics is described<br />

by the SDOF equation of motion as<br />

follows:<br />

Zt ̇̇ Zt ̇ 2 Ft ()<br />

() + 2ξω<br />

s<br />

() + ω<br />

s<br />

Zt () = (1)<br />

M<br />

in which Z(t) is the generalized coordinate<br />

expressing the motion of the system,<br />

t, the time, being the independent<br />

variable; z is the damping ratio; w s =<br />

2πf s is the natural circular frequency of<br />

the mode of interest; F(t) and M are<br />

the generalised force and mass, respectively,<br />

expressed as follows:<br />

Ft () = ft () ∫ ϕ ( x)<br />

d x<br />

L<br />

0<br />

L<br />

(2)<br />

M= m∫ ϕ( x)2<br />

d x<br />

(3)<br />

0<br />

where L is the length of the footbridge<br />

deck, x is the space coordinate along<br />

the longitudinal axis of the footbridge,<br />

j (x) is the mode shape, f(t) and m are<br />

the force and mass per unit length,<br />

which are constant in space according<br />

to hypotheses 2 and 3.<br />

The mass m is intended to be the overall<br />

mass of the structure and crowd<br />

systems:<br />

m= m + m ( ρ) , withm ( ρ)<br />

= ρ BG (4)<br />

s p p<br />

where the subscripts s and p refer to<br />

the structure and pedestrians, respectively,<br />

r (ped/m 2 ) is the crowd density,<br />

G = 70 kg is the average mass of one<br />

pedestrian and B is the width of the<br />

walking path.<br />

The lateral force f(t) per unit length<br />

is expressed on the basis of the load<br />

model proposed by Venuti et al. 10 and<br />

Venuti and Bruno. 14 The original force<br />

model had been conceived to account<br />

for both pp and ps synchronisation<br />

and to describe the crowd force during<br />

the triggering, self-exciting and selflimiting<br />

SLE stages. According to the<br />

above-mentioned simplifying assumptions,<br />

the lateral force model reduces<br />

to the following:<br />

f() t = ρeq<br />

B⋅F0 sin( ωpt)<br />

(5)<br />

with<br />

ρ eq<br />

= N eq<br />

/( BL) (6)<br />

and<br />

N = N S + N( 1 −S<br />

) (7)<br />

eq pp pp<br />

where r eq is the equivalent crowd density;<br />

F 0 = 28 N is the amplitude of the<br />

lateral force exerted by one pedestrian;<br />

w p = 2πf p is the circular lateral walking<br />

frequency; N = rBL is the number<br />

of pedestrians on the deck; S pp is the<br />

coefficient of synchronisation among<br />

pedestrians. The square root term in<br />

Eq. (7) represents the contribution<br />

of uncorrelated pedestrians, according<br />

to the model of Matsumoto et al. 15<br />

The lateral walking frequency f p is<br />

expressed as a function of the walking<br />

velocity v, in turn dependent on<br />

the crowd density, according to the<br />

following laws:<br />

2 3<br />

fp = ( 293 , v − 159 , v + 035 , v )/<br />

2<br />

⎧⎪<br />

⎡ ⎛ 1 1 ⎞ ⎤⎫⎪<br />

v= vM<br />

⎨1<br />

− exp ⎢−γ<br />

−<br />

⎝<br />

⎜<br />

⎠<br />

⎟ ⎥⎬<br />

⎩⎪ ⎣ ρ ρM<br />

⎦⎭⎪<br />

(8)<br />

(9)<br />

where the values of the coefficient g,<br />

the free speed v M and the jam density<br />

r M are made sensitive to different travel<br />

purposes (L = leisure/shopping, C = commuters/events,<br />

R = rush hour/business)<br />

and geographic areas (E = Europe,<br />

U = USA, A = Asia) (Fig. 1). The laws<br />

in Eqs. (8) and (9), first proposed by<br />

Venuti and Bruno, 16 are fitted to the<br />

experimental data reported in Refs.<br />

[5, 17, 18]. The values of the empirical<br />

parameters g, n M , r M of interest for the<br />

case studies discussed in the following<br />

section are reported in Table 2.<br />

As far as the synchronisation coefficient<br />

S pp is concerned, in the last 2<br />

years some new studies 5,19,20 concerning<br />

pp synchronisation due to crowd<br />

density have been carried out by means<br />

of experimental tests performed within<br />

different ranges of crowd density.<br />

Araújo et al. 19 found that, in the crowd<br />

density range from 0,3 to 0,9 ped/m 2 ,<br />

there is no evidence of synchronisation<br />

among pedestrians, since the standard<br />

deviation of the walking frequencies<br />

is almost constant for different densities<br />

and the phase angles are totally<br />

random. Ricciardelli and Pansera 20<br />

observed that, in the crowd density<br />

range 0,5 to 1,5 ped/m 2 , initially different<br />

walking frequencies and phases<br />

tend to get closer for increasing crowd<br />

densities, giving rise to synchronisation<br />

nuclei within the crowd. Finally,<br />

Butz et al., 5 who performed tests with<br />

the highest values of the crowd density<br />

(1,2–3,0 ped/m 2 ), observed a reduction<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 443


v (m/s)<br />

(a)<br />

2<br />

1,5<br />

1<br />

0,5<br />

0<br />

0<br />

Venuti and Bruno (2007) - L<br />

-C<br />

2,5<br />

-R<br />

Butz et al. (2008)<br />

2<br />

Oeding (1963) - L<br />

-C<br />

1,5<br />

-R<br />

1<br />

Bertram and Ruina (2001)<br />

Venuti et al. (2007)<br />

0,5<br />

Butz et al. (2008)<br />

Butz et al. (2008)<br />

0<br />

1 2 3 4 5 6 0 0,5 1 1,5 2 2,5<br />

r (ped/m 2 )<br />

(b)<br />

v (m/s)<br />

Fig. 1: v(r) (a) and f p (v) (b) laws and comparison with the ones proposed in Ref. [5]<br />

of the standard deviation of the walking<br />

frequencies for low walking speed,<br />

that is, for increasing density, but they<br />

did not report any data concerning<br />

the phase angles. Owing to the scarceness<br />

of quantitative results concerning<br />

the phase angles, in the following,<br />

the standard deviation of the walking<br />

frequency s f will be considered as a<br />

suitable indicator of synchronisation,<br />

that is, the smaller s f is, the higher the<br />

pedestrian tendency to walk at the<br />

same mean walking frequency (Fig. 2).<br />

On the basis of these experimental<br />

observations, the qualitative S pp (r)<br />

law proposed in Ref. [10] is replaced<br />

by a new law developed according to<br />

the scheme represented in Fig. 3a as<br />

complementary to s f /s f,0 , where s f,0 =<br />

0,12 Hz is the mean standard deviation<br />

in the case of unimpeded walking<br />

measured in the same experimental<br />

s f (Hz)<br />

0,12<br />

0,08<br />

0,04<br />

Butz et al. (2008)<br />

Araujo et al. (2009)<br />

0 0 0,5 1 1,5 2<br />

r (ped/m 2 )<br />

Fig. 2: Values of the standard deviation s f as<br />

measured in Refs. [5, 19]<br />

1<br />

s f /s f,0<br />

S pp<br />

2f p (Hz)<br />

campaigns. 5,19 The fitting function<br />

is inspired to the trend of the coherence<br />

against the coupling strength in<br />

the Kuramoto model 21 : the coherence<br />

(i.e. S pp herein) is null until the coupling<br />

strength (i.e. r herein) reaches<br />

a threshold value (r c ), which corresponds<br />

to a phase transition; for r > rc<br />

the coherance grows towards perfect<br />

synchronisation. S pp (r) is, therefore,<br />

expressed as follows:<br />

S<br />

0<br />

c<br />

pp( )<br />

,<br />

1 exp[ a(<br />

c)/<br />

M<br />

]<br />

c<br />

(10)<br />

where a = 8,686 derives from the fitting<br />

to the set of points of coordinates<br />

(r, 1 − s f /s f,0 ) and r c is set equal to<br />

0,6 ped/m 2 (Fig. 3b). Figure 4 plots<br />

the trend of r eq /r M versus r/r M (the<br />

trend of F is proportional): the function<br />

is bounded between two limit<br />

curves, which represent the cases of<br />

all pedestrians synchronised to each<br />

other ( ρ eq,sync<br />

= NBL, / e.g. marching<br />

soldiers) and all pedestrians uncorrelated<br />

( ρ eq,rand<br />

= NBL / ), respectively.<br />

The proposed law matches the<br />

model of uncorrelated pedestrians for<br />

r < r c , while it tends to the full synchronisation<br />

as r approaches r M . It<br />

is worth pointing out that the density<br />

range in which the pp synchronisation<br />

develops and the law differs from the<br />

limit curves corresponds to the one of<br />

0,2<br />

Butz et al.’s data<br />

Araujo et al.’s data<br />

0 0<br />

Fitting<br />

0 r c /r M<br />

1<br />

0 r 0,2 0,4 0,6 0,8<br />

r/r c<br />

M<br />

(a)<br />

(b) r M<br />

r/r M<br />

Fig. 3: S pp (r) relation: scheme (a) and proposed law (b)<br />

S pp<br />

1<br />

0,8<br />

0,6<br />

0,4<br />

1−s f /s f,0 from:<br />

1<br />

r eq /r M<br />

10 0<br />

10 −1<br />

10 −2<br />

10 −3 0 r c<br />

r M<br />

practical interest in real world crowd<br />

events on footbridges.<br />

The amplitude of the steady-state lateral<br />

acceleration can be found referring<br />

to the following well-known<br />

expression:<br />

Ż̇<br />

F M D ρeqBF<br />

= =<br />

m<br />

0,2 0,4 0,6 0,8 1<br />

∫<br />

∫<br />

L<br />

0 0<br />

L<br />

0<br />

r/rM<br />

Fig. 4: Diagram of r eq versus r<br />

r eq<br />

r eq,sync<br />

r eq,rand<br />

ϕ( x)<br />

dx<br />

D<br />

2<br />

ϕ( x)<br />

dx<br />

(11)<br />

where F is the amplitude of<br />

the generalised force and<br />

2 2 2 −05<br />

D= [( 1− f ) + ( 2 f ) ] .<br />

r<br />

ξ<br />

r<br />

is the<br />

dynamic amplification factor, f r = f p /f s<br />

being the frequency ratio. Substituting<br />

Eqs. (4)–(10) in Eq. (11), the amplitude<br />

of the steady-state acceleration<br />

can be expressed as a function of the<br />

crowd density r through the variables<br />

r eq , f r and m. For instance, Fig. 5 plots<br />

the trend of Z ̇̇ versus r for four different<br />

combinations of geographic<br />

area (Europe E and Asia A) and<br />

travel purpose (rush hour R and leisure<br />

L) and for given structural properties<br />

of the footbridge (L = 90 m, B<br />

= 4 m, m s = 2000 kg/m, x = 0,005, f s<br />

= 0,9 Hz). It should be stressed that<br />

the four curves are valid for Ż̇<br />

≤ ż̇<br />

lim<br />

,<br />

while the grey branches represent<br />

the ideal trend of Z ̇̇ in the absence<br />

of ps synchronisation. The blue dots<br />

indicate the limit condition (r lim , ̇ż lim<br />

).<br />

The travel purpose especially affects<br />

the value of the structural response,<br />

by modifying the walking frequency–<br />

density relation and the f r values in<br />

turn. The geographic area parameter<br />

has the main effect of varying the<br />

value of the maximum density r M ,<br />

therefore shifting the maximum structural<br />

response.<br />

It is worth pointing out that Z ̇̇ = Z ̇̇ ( ρ )<br />

is not a bijective function, as clearly<br />

shown in Fig. 5; therefore, Eq. (11)<br />

is not invertible and the value of<br />

r lim is herein determined through<br />

an iterative procedure based on<br />

the “Goal Seek” tool in Microsoft ®<br />

Excel ® , even if other algorithms can<br />

444 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


z lim (m/s 2 )<br />

..<br />

be programmed and employed (e.g.<br />

the ones based on dicothomic search).<br />

The obtained critical density value<br />

can be easily checked by substituting<br />

it in Eq. (11) and verifying that the<br />

corresponding acceleration amplitude<br />

is close to but lower than the<br />

perceptibility threshold ̇ż lim<br />

.<br />

Application and Comparison<br />

with Other Criteria<br />

The proposed approach is applied to<br />

three case studies and the results are<br />

compared to those obtained with other<br />

serviceability and stability criteria.<br />

The first comparison is performed by<br />

applying the serviceability criterion<br />

presented herein, but substituting the<br />

accompanying force model with the<br />

one proposed by Sétra/AFGC 4 :<br />

f() t = N F sin( ω t)<br />

N<br />

eq<br />

10 1<br />

10 0<br />

..<br />

z lim<br />

10 −2<br />

eq<br />

0<br />

s<br />

2<br />

⎧⎪<br />

10, 8 Nξ<br />

ρ<<br />

1 ped/m (12)<br />

= ⎨<br />

⎩⎪ 185 , N ρ ≥ 1ped/m 2<br />

with F 0 = 35 N. The equivalent number<br />

of pedestrians N eq has been<br />

N lim =<br />

10 −3 10 −1 10 0<br />

Dallard<br />

et al. 8 Newland 12 Roberts 13<br />

8πξ fM s<br />

k<br />

s<br />

r (ped/m 2 )<br />

2ξmL<br />

s<br />

αβG<br />

E-R<br />

E-L<br />

A-R<br />

A-L<br />

Fig. 5: Diagram of ̇̇ Z versus r for different<br />

combinations of geographic area and<br />

travel purpose<br />

mL<br />

s<br />

2<br />

Gf D<br />

Table 1: Stability criteria proposed in Refs.<br />

[8, 12, 13]<br />

r<br />

10 1<br />

obtained as the number of pedestrians<br />

—uniformly distributed and walking in<br />

phase with the same frequency as the<br />

footbridge—that produces the same<br />

effect as random pedestrians.<br />

Three stability criteria are applied and<br />

shortly summarised in the following<br />

(Table 1). In the equation proposed by<br />

Dallard et al., 8 M s is the modal mass<br />

of the bridge and k = 300 N s/m is a<br />

proportionality constant that has been<br />

tuned on the London Millennium<br />

Bridge experimental data. The a and<br />

b coefficients in Newland’s equation<br />

12 represent the ratio between the<br />

motion of the pedestrian centre of<br />

mass and the bridge motion (equal to<br />

2/3) and the percentage of pedestrians<br />

synchronised to the structure (equal to<br />

0,4), respectively. As for the product<br />

f 2 r D in the Roberts’ equation, 13 mean<br />

values are suggested by the author for<br />

different ranges of the frequency ratio<br />

f r . Differently from the two previous<br />

models, the last one does not necessarily<br />

imply that f p = f s .<br />

The case studies chosen for the applications<br />

are the T-bridge (T-br.) and<br />

M-bridge (M-br.) in Japan 7,22 and the<br />

south span of the London Millennium<br />

Bridge (LM br.). 8 The input data<br />

adopted for the calculations are summarised<br />

in Table 2, while the results<br />

in terms of r lim are compared in<br />

Fig. 6. For each case study, the results<br />

obtained with both the stability (black<br />

points) and the serviceability criteria<br />

(red points) are reported. In particular,<br />

the present results distinguish<br />

among different travel purposes and<br />

the filled points refer to the expected<br />

traffic condition in service. First, it can<br />

be observed that the values of r lim<br />

calculated with the serviceability criteria<br />

are generally higher than those<br />

derived through the stability criteria.<br />

This could be explained by considering<br />

that the force models adopted to<br />

derive stability criteria usually assume<br />

the force as proportional in amplitude<br />

and resonant to the deck motion,<br />

thus inducing a higher response. In<br />

particular, the model of Dallard et al.<br />

provides by far the lowest values of<br />

r lim ; moreover, this model has been<br />

specifically tuned on the LM br. data,<br />

therefore, its applicability to other<br />

case studies is questionable. It can be<br />

observed that the criteria of Sétra,<br />

Newland and Roberts result in comparable<br />

values of the limit density in<br />

all the case studies, even though the<br />

criteria rely on rather different modelling<br />

assumptions. The method proposed<br />

in this article leads to the less<br />

conservative results for the LM br.<br />

and M-br. cases, allowing a denser<br />

crowd on the bridge to reach the limit<br />

acceleration. It should be pointed out<br />

that, for each case study, the lowest<br />

r lim value is associated to the travel<br />

purpose, which determines the f r value<br />

closest to unit: in other words, when f r<br />

is near the unit, the proposed method<br />

provides results that are close to those<br />

obtained with the criteria that assume<br />

resonant conditions. Generally speaking,<br />

the lateral force is not applied in<br />

resonance with the deck motion, but<br />

the walking frequency is determined<br />

as a function of the crowd density<br />

and is affected by the travel purpose.<br />

Hence, the present model allows the<br />

designer and the owner to take into<br />

account different traffic scenarios and<br />

retain the most suitable one: the most<br />

recurrent scenario, if it can be set (e.g.<br />

a footbridge linking a bus terminal, as<br />

in the T-br. case) or the one that gives<br />

the most conservative result (i.e. the<br />

lowest r lim value).<br />

Conclusion<br />

In this paper, a simplified criterion for<br />

the assessment of footbridge serviceability<br />

under lateral crowd loading has<br />

been presented. It is based on the derivation<br />

of the limit crowd density that<br />

induces a lateral acceleration of the<br />

deck, corresponding to the threshold<br />

of human perceptibility. The approach<br />

relies on a load model that accounts<br />

not only for the action of random<br />

pedestrians but also for the possibility<br />

of pp synchronisation due to crowd<br />

density. Moreover, the proposed lateral<br />

load model describes the actual walking<br />

behaviour of pedestrians by referring<br />

to constitutive laws, namely, the<br />

speed–density and frequency–speed<br />

relations.<br />

L (m) B (m) m s (kg/m) f s (Hz) w M s (kg) f R–C–L q M (ped/m 2 ) v M (m/s) R–C–L f 2 r D<br />

T-br. (A) 179 5,25 4200 0,93 0,0113 214 010 2,1–1,65–1,89 7,7 1,48–1,37–1,04 14<br />

LM br. (E) 108 4 2000 0,8 0,007 160 000 1,64–1,28–1,47 6 1,69–1,56–1,12 16,2<br />

M-br. (A) 320 1,5 600 1,025 0,0027 97 200 2,1–1,65–1,89 7,7 1,48–1,37–1,04 28<br />

Table 2: Input data for the three case studies: T-bridge, London Millennium bridge (south span), M-bridge (central span)<br />

Structural Engineering International 4/<strong>2010</strong> Scientific Paper 445


lim (ped/m 2 )<br />

1,4 L<br />

1,2<br />

1<br />

R<br />

0,8<br />

C<br />

0,6<br />

0,4<br />

0,2<br />

0<br />

R<br />

C<br />

L<br />

T-br. LM br. M-br.<br />

Fig. 6: Comparison between the r lim (ped/m 2 )<br />

obtained with the different approaches<br />

L<br />

C<br />

R<br />

R - C - L<br />

Sétra/AFGC<br />

Dallard et al.<br />

Newland<br />

Roberts<br />

With respect to stability criteria,<br />

serviceability criteria appear more<br />

appropriate for preventing the occurrence<br />

of SLE, since they are based on<br />

the description of pedestrian behaviour<br />

in the pre-lock-in stage. The application<br />

to real case studies has shown<br />

that stability criteria are generally<br />

more conservative and imply more<br />

severe restrictions to the footbridge<br />

service than the serviceability criteria.<br />

According to the authors, the scattered<br />

results coming from the applications<br />

of the criteria are mainly due to the<br />

rather different load models that the<br />

criteria rely on and they highlight the<br />

fact that the definition of a universally<br />

accepted criterion to prevent SLE is<br />

still an open issue. In particular, the<br />

mechanism of synchronisation among<br />

pedestrians due to crowd density<br />

requires further experimental tests to<br />

be properly measured and completely<br />

clarified.<br />

To the authors’ opinion, the proposed<br />

approach and load model offer<br />

a framework susceptible to be easily<br />

adapted in the case of vertical crowd<br />

load and resulting vibrations.<br />

References<br />

[1] Živanovic S, Pavic A, Reynolds P. Vibration<br />

serviceability of footbridges under humaninduced<br />

excitation: a literature review. J. Sound<br />

Vib. 2005; 279: 1–74.<br />

[2] Venuti F, Bruno L. Crowd-structure interaction<br />

in lively footbridges under synchronous<br />

lateral excitation: a literature review, Phys. Life<br />

Rev. 2009; 6: 176–206.<br />

[3] Federation International du Beton (FIB).<br />

Guidelines for the Design of Footbridges, Bulletin<br />

No. 32, Lausanne, 2006.<br />

[4] Service d’Études Techniques des Routes<br />

et Autoroutes (Sétra/AFGC). Passerelles<br />

piétonnes – Evaluation du comportement<br />

vibratoire sous l’action de piétons. Guide<br />

méthodologique. Paris, 2006.<br />

[5] Butz C, Feldmann M, Heinemeyer C, Sedlacek<br />

G, Chabrolin B, Lemaire A, et al. Advanced load<br />

models for synchronous pedestrian excitation and<br />

optimised design guidelines for steel footbridges<br />

(SYNPEX), Report RFS-CR 03019, Research<br />

Fund for Coal and Steel, 2008.<br />

[6] HIVOSS, Design of Footbridges: Guideline<br />

and Background Documents, Research Fund for<br />

Coal and Steel, 2008.<br />

[7] Fujino Y, Pacheco BM, Nakamura S,<br />

Warnitchai P. Synchronisation of human walking<br />

observed during lateral vibration of a congested<br />

pedestrian bridge. Earthquake Eng. Struct. Dyn.<br />

1993; 22: 741–758.<br />

[8] Dallard P, Fitzpatrick T, Le Bourva S,<br />

Low A, Ridsdill RM, Willford M. The London<br />

Millennium Footbridge. Struct. Eng. 2001; 79(22):<br />

17–33.<br />

[9] Ricciardelli F. Lateral loading of footbridges<br />

by walkers. Proceedings Footbridge 2005, Venice,<br />

2005.<br />

[10] Venuti F, Bruno L, Napoli P. Pedestrian<br />

lateral action on lively footbridges: a new load<br />

model. Struct. Eng. Int. 2007; 17(3): 236–241.<br />

[11] Brownjohn JMW, Živanovic S, Pavic A.<br />

Crowd dynamic loading on footbridges.<br />

Proceedings Footbridge 2008, Porto, 2008.<br />

[12] Newland DE. Pedestrian excitation of<br />

bridges. Proceedings of the institution of<br />

mechanical engineers. J. Mech. Eng. Sci. 2004;<br />

218c: 477–492.<br />

[13] Roberts TM. Lateral pedestrian excitation<br />

of footbridges. J. Bridge Eng. 2005; 10:<br />

107–112.<br />

[14] Bruno L, Venuti F. Crowd-structure interaction<br />

in footbridges: modelling, application to a<br />

real case-study and sensitivity analyses. J. Sound<br />

Vib. 2009; 323: 475–493.<br />

[15] Matsumoto Y, Nishioka T, Shiojiri H,<br />

Matsuzaki K. Dynamic design of footbridges.<br />

IABSE Proc. 1978; P17/78: 1–15.<br />

[16] Venuti F, Bruno L. An interpretative model<br />

of the pedestrian fundamental relation. C. R. Mec.<br />

2007; 335: 194–200.<br />

[17] Oeding D. Verkehrsbelastung und<br />

Dimensionierung von Gehwegen und anderen<br />

Anlagen des Fußgängerverkehrs. Straßenbau<br />

und Straßenverkehrstechnik 1963; 22.<br />

[18] Bertram JE, Ruina A. Multiple walking<br />

speed-frequency relations are predicted by constrained<br />

optimization. J. Theoret. Biol. 2001; 209:<br />

445–453.<br />

[19] Araújo MC, Brito HMBF, Pimentel RL.<br />

Experimental evaluation of synchronisation in<br />

footbridges due to crowd density. Struct. Eng.<br />

Int. 2009; 19(3): 298–303.<br />

[20] Ricciardelli F, Pansera A. An experimental<br />

investigation into the interaction among walkers<br />

in groups and crowds. 10th International<br />

Conference on Recent Advances in Structural<br />

Dynamics RASD <strong>2010</strong>, Southampton.<br />

[21] Strogatz SH. From Kuramoto to Crawford:<br />

exploring the onset of synchronization in<br />

populations of coupled oscillators. Phys. D 2000;<br />

143: 1–20.<br />

[22] Nakamura S, Kawasaki T. A method for predicting<br />

the lateral girder response of footbridges<br />

induced by pedestrians. J Constr. Steel Res. 2009;<br />

65: 1705–1711.<br />

446 Scientific Paper Structural Engineering International 4/<strong>2010</strong>


The Construction of the Main Bridge of the Yichang<br />

Yangtze River Railway Bridge in China<br />

Yiqiao Zhou, Civil Eng.; Lichao Zhang, Civil Eng.; China Railway Major Bridge Engineering Co. Ltd., Wuhan, China.<br />

Contact: yqzhou@ztmbec.com<br />

Abstract<br />

Railway construction technology in<br />

China, especially of high-speed railway<br />

long-span bridge construction, has been<br />

developing rapidly since the beginning<br />

of the twenty-first century. To maintain<br />

the dynamic characteristics, travel<br />

safety and passenger comfort when<br />

the train travels over a bridge at high<br />

speed, and, at the same time, to meet<br />

the requirement of economical and<br />

technical viability of construction of the<br />

bridge—was a challenging problem. For<br />

solving this problem, many new types<br />

of bridge structure have been developed<br />

for railway and high-speed railway<br />

long-span bridges. The arch beam<br />

hybrid structure is one of the structures<br />

that have been widely used in China. It<br />

is normally made up of a prestressed<br />

concrete (PC) beam, a rigid frame, or<br />

V-shaped piers and a steel arch or concrete-filled<br />

steel tube arch.<br />

Keywords: hybrid structure; long-span<br />

PC rigid frame; concrete-filled steel<br />

tube flexible arch; arch rib rotate–<br />

lifting synchronously; railway bridge.<br />

Introduction<br />

Recently, a long-span arch beam hybrid<br />

bridge structure has been successfully<br />

used in constructing the main structure<br />

of the Yichang Yangtze River Railway<br />

Bridge, which combined the PC rigidframe<br />

girder with a concrete-filled steel<br />

tube flexible arch. The bridge carries<br />

double rail lines designed to support<br />

a train velocity of 160 km/h. The span<br />

arrangement of the main bridge is 130<br />

+ 2 × 275 + 130 m. This is not only the<br />

first use of this kind of structure for a<br />

railway bridge in the world but also,<br />

with a main span length of 275 m, the<br />

bridge ranks among the longest of its<br />

type in the world at present.<br />

General Description<br />

This bridge across the Yangtze River<br />

at Yichang City is a vital link between<br />

the newly constructed railway line in<br />

Yichang, Hubei Province and the city<br />

of Wanzhou in the Chongqing municipality.<br />

The span arrangement of whole<br />

bridge from the southern to the northern<br />

bank is 10 × 49,2 m simply supported<br />

PC box girders + (130 + 2 × 275<br />

+ 130 m) continuous PC rigid-frame<br />

girder with concrete-filled steel tube<br />

flexible arch + 14 × 48,2 m simply supported<br />

PC box girder + (56 + 108 +<br />

56 m) continuous PC girder + 9 × 32 m<br />

simply supported PC beam; the total<br />

length is 2526,73 m. A three-directional<br />

pre-stress system and C60 highperformance<br />

concrete was adopted<br />

for the PC girder of the main bridge.<br />

The structure of arch rib is a parallel<br />

concrete-filled steel tube truss and the<br />

arch axis is a quadratic parabola. The<br />

calculated span of the arch is 264 m<br />

and the arch rise is 52,8 m. The rise–<br />

span ratio is 1/5,0. C50 grade microexpansion<br />

concrete has been filled in<br />

the rib tube, and the suspenders that<br />

connect the arch rib and PC girder are<br />

parallel steel strands (Fig. 1).<br />

Construction of Substructure<br />

The foundation of the main bridge<br />

comprises reinforced bored piles; all<br />

the pile tips were embedded into the<br />

solid bedrock beneath the riverbed.<br />

There are 12 piles (each of diameter<br />

Φ 3,0 m) for each of the three main<br />

piers and 9 piles (Φ 2,0 m) for each<br />

of the two side piers. The pile caps are<br />

17,0 m × 23,0 m × 5,0 m rectangular<br />

reinforced concrete structures with<br />

rounded corners in cross section. A<br />

rectangular single-cell reinforced concrete<br />

pier shaft of cross section 8,0 m ×<br />

12,0 m has been used for Pier 12, and<br />

a twin-wall reinforced concrete pier<br />

shaft of cross section 3,0 × 12,0 m, with<br />

5,0 m central spacing between walls is<br />

adopted for Piers 11 and 13. The height<br />

of main pier shaft is 38,5 m. Each of<br />

the two side piers (Piers 10 and 14)<br />

has a solid rectangular shaft with cutcornered<br />

cross section and a tray-type<br />

pier cap.<br />

Piling works were carried out from<br />

the working platform over water by<br />

a boring machine using the air-lifting<br />

reversed-circulation method. When<br />

all the piles of a pier had been constructed,<br />

a steel cofferdam was sunk<br />

to the designed depth. Construction<br />

of the pile cap and the lower section<br />

of pier shaft commenced after the bottom<br />

was sealed by tremie concrete and<br />

the water was pumped out of the cofferdam.<br />

The remaining sections of the<br />

pier shaft were constructed by means<br />

of a climbing formwork.<br />

Construction of PC<br />

Rigid-Frame Girder<br />

The superstructure of the main bridge<br />

is a single-box double-cell PC girder<br />

130,8<br />

275,0 275,0 130,8<br />

Yichang<br />

Wanzhou<br />

37,0<br />

26,0 5,0<br />

45,0 5,0<br />

40,0 5,0<br />

4,0<br />

30,0<br />

38,5<br />

38,5<br />

38,5<br />

44,0<br />

14,5<br />

14,5<br />

14,5<br />

10<br />

9φ1,5 m bored piles<br />

11<br />

12φ3,0 m bored piles<br />

12<br />

12φ3,0 m bored piles<br />

12φ3,0 m bored piles<br />

13<br />

9φ2,0 m bored piles<br />

34,0 4,0<br />

14<br />

Fig. 1: Elevation for the main bridge (Units: m)<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 447


with trapezoidal side webs. The width<br />

of the top flange is 14,4 m and that of<br />

the bottom flange varies from 9,2 m at<br />

the pier top to 12,7 m at mid span. The<br />

girder depth is 14,5 m at pier top and<br />

4,8 m at mid span. The thickness of the<br />

top flange is 500 and 400 mm for double-<br />

and single-layer tendon arrangement,<br />

respectively. The thickness of the<br />

web is changed gradually from 600 mm<br />

at the pier top to 300 mm at mid span;<br />

the thickness of the bottom flange varies<br />

from 1400 mm near the pier top to<br />

350 mm at mid span.<br />

The pier top segments were constructed<br />

on brackets that were installed<br />

on the pier heads. To ensure construction<br />

quality of the pier top segment,<br />

the concreting-works was divided<br />

into two pours horizontally. The first<br />

pouring height was 8,7 m, including<br />

the 1,2 m pier top shaft; the second<br />

one was 7,0 m up to the top flange.<br />

The balanced cantilever construction<br />

method was used for constructing the<br />

girder and there were 177 segments<br />

in all (Fig. 2). The longitudinal length<br />

of the segments was divided into 3,0,<br />

3,5, 4,0, 4,5 and 5,0 m, and the heaviest<br />

segment was 3583 kN. Travelling<br />

formwork was composed of the main<br />

truss, travelling system, front and rear<br />

anchorage system, formwork and<br />

hanging system. The end segments of<br />

the side spans were concreted on the<br />

side pier tops and brackets.<br />

Galvanized steel strands (Φ 15,24 mm)<br />

were adopted for the longitudinal and<br />

transverse prestressing systems. 128<br />

longitudinal tendons and 31 strands<br />

for each were arranged in the top<br />

flange at the main piers; they were<br />

used to meet the requirement for<br />

the cantilevering construction of the<br />

girder segments. 28 longitudinal tendons<br />

and 19 strands for each were<br />

installed in the bottom slab of the side<br />

spans—these were tensioned after<br />

closure of the girder. Considering<br />

that the girder should resist the thrust<br />

force from the arch, 12 and 40 tendons,<br />

each consisting of 19 strands,<br />

were arranged in the top and bottom<br />

flange, respectively, at midway across<br />

the main central spans. Transverse<br />

prestressing tendons were arranged in<br />

the top flange only, with five strands<br />

for each tendon and longitudinal spacing<br />

of 800 mm. A Prestressing Screw<br />

Bar (PSB) 930 (Φ 32 mm) threaded<br />

steel bar was embedded in the webs<br />

as the vertical pre-stressing system<br />

and the longitudinal spacing was<br />

400 mm. For ensuring grouting quality,<br />

the vacuum auxiliary grouting<br />

method was applied for all tendon<br />

and threaded steel bar ducts for elimination<br />

of the air content in the grout<br />

and to make the grout completely fill<br />

the ducts.<br />

The PC frame girder followed the closure<br />

sequence of the side spans first,<br />

followed by the central spans. The two<br />

side spans were allowed closure at different<br />

times, but the closure of the two<br />

central spans had to occur simultaneously.<br />

The practical closure elevation<br />

difference is 8 and 4 mm on the two<br />

side spans, 3 and 6 mm; for the two<br />

main spans respectively, these were<br />

much lower than the allowable error<br />

according to the industrial specification<br />

in China. 1<br />

To reduce the influence of concrete<br />

shrinkage and creep, and enhance the<br />

cracking resistance ability of the concrete,<br />

C60 high-performance concrete<br />

was used for the girder structure. A<br />

compound–mineral admixture (namely<br />

ground slag and fly ash mixture) was<br />

used for the concrete to enhance its<br />

compactness and crack resistance<br />

ability, and reduce hydration heat in<br />

the early stages. In addition, a polycarboxylic<br />

acid water-reducing agent<br />

was introduced to optimize the mixture<br />

ratio, reduce the slump loss, and<br />

enhance the fluidity of the concrete.<br />

Installation of Steel Tube Arch<br />

The main arch is composed of two arch<br />

ribs and the central spacing is 12,35 m.<br />

Each arch rib is composed of four steel<br />

tubes (Φ 750 mm) that are arranged in<br />

a double dumb-bell-shaped cross section.<br />

Steel tubes (Φ 450 mm) connect<br />

upper and lower Φ 750 mm tubes<br />

vertically to form a truss plan, and a<br />

steel plate connects the left and right<br />

Φ 750 mm tubes transversely. The central<br />

spacing of the four steel tubes is<br />

1,7 m in the transverse direction, 4,0 m<br />

at the arch toe and 3,0 m at its crown<br />

in the vertical direction. The two arch<br />

ribs are connected by 11 lateral braces<br />

on each span (Fig. 3).<br />

1700<br />

φ = 750<br />

φ = 450<br />

Fig. 2: Construction of girder<br />

Fig. 3: Typical cross section of an arch<br />

rib (Units: mm)<br />

448 Technical Report Structural Engineering International 4/<strong>2010</strong>


The arch rib was manufactured in sections<br />

in a factory, and delivered at the<br />

site in barges. In order to assemble and<br />

install the main arch, arch rib assembling<br />

falsework on the bridge deck and<br />

three stay cable towers on the top of<br />

each main pier were installed. The sections<br />

of the arch rib were lifted from<br />

the barge and placed on the falsework<br />

by cranes (Fig. 4).<br />

After the four half-span arch ribs had<br />

been assembled, they were rotated<br />

vertically and closured by the lifting<br />

system, which includes stay cable towers,<br />

balance cables, stay cables, anchor<br />

cables, anchorages and hydraulic jacks.<br />

The magnitude and distribution of the<br />

stay forces were the critical factors for<br />

successful rotation and closure of the<br />

ribs. Each stay cable was controlled<br />

by a hydraulic jack with lifting capacity<br />

of 3200 kN. All the hydraulic jacks<br />

were operated by a computer that<br />

controlled the cable force and ensured<br />

Fig. 4: Assembling the arch rib<br />

Fig. 5: Lifting and rotating the arch rib<br />

synchronized working of the jacks during<br />

operation. The main arch between<br />

Piers 11 and 12 was rotated and closured<br />

first, and closure of the main<br />

arch of Piers 12 and 13 followed thereafter<br />

(Fig. 5).<br />

Once the arch ribs of two spans were<br />

closured and the lateral bracing members<br />

between the two main arches<br />

installed, C50 micro-expansion concrete<br />

was filled in the rib tubes. Since<br />

normal concrete has autogenous<br />

shrinkage and temperature shrinkage<br />

properties, the concrete in the tube<br />

could separate from the inner wall of<br />

the steel tube, thus decreasing the loadbearing<br />

capacity of the concrete-filled<br />

steel tube. This meant that the concrete<br />

filled into the rib tube should have selfcompacting<br />

and self-expanding properties.<br />

In order to meet the requirement,<br />

the concrete mix ratio was carefully<br />

selected and vacuum auxiliary concrete-pumping<br />

method was adopted.<br />

Concrete was filled in the Φ 750 mm<br />

steel tube of the arch ribs and the space<br />

between the upper and lower lateral<br />

connecting steel plates, the Φ 450 mm<br />

steel tube was kept empty. Concrete<br />

filling was carried out synchronously<br />

both up- and downstream of the arch<br />

rib tubes of the same span. Concrete<br />

was pumped up from the toes of the<br />

rib tubes; the grout let-out valves and<br />

vacuumizing vents were installed on<br />

the crown of the rib tubes. The vacuum<br />

level in the tubes was kept between<br />

−0,1 and −0,09 MPa.<br />

Every suspender connecting the arch<br />

rib and PC girder is composed of a<br />

pair of cables that are composed of<br />

Φ 15,24 mm galvanized steel strands.<br />

The longitudinal interval between<br />

suspenders is 10 m and there are 100<br />

suspenders (200 strand cables) in<br />

total. Each cable consists of 15 strands,<br />

except for the three pairs that are near<br />

each arch toe, which consist of nine<br />

strands. The upper end of suspender<br />

was the tensioning end and anchored<br />

in the rib tubes; the lower end was the<br />

fixed end and anchored in the stiffened<br />

cross-beam beneath the bridge deck.<br />

A dynamometer was installed on the<br />

fixed end of every cable in order<br />

to monitor cable force during the<br />

construction and operation stages<br />

to ensure safety of the structure. All<br />

the strands were wax coated and<br />

Polyethylene (PE) sheathed individually<br />

for corrosion protection. High<br />

Density Polyethylene (HDPE) duct<br />

was applied for the external sheath of<br />

all cables.<br />

Since the Yichang Yangtze River<br />

Railway Bridge is a very complex<br />

hybrid structure, the load borne by the<br />

structure in every construction stage<br />

is different, and hence structure alignment<br />

control and stress measurement<br />

were very important during the girder<br />

cantilever construction and arch rib<br />

installation; the structure alignment and<br />

stress monitoring and control works<br />

was therefore carried out from the initial<br />

till the final stages of construction.<br />

The measurement points were set out<br />

and strain gauges were embedded in<br />

the structure; the calculated stress and<br />

displacement data were used to guide<br />

construction work in all stages.<br />

On completion of construction of<br />

the Yichang Yangtze River Railway<br />

Bridge, the dead and dynamic load<br />

tests were carried out on the bridge.<br />

The test results show that all the design<br />

requirements of the bridge structure<br />

have been fulfilled (Fig. 6).<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 449


the arch and the girder. The whole<br />

structure thus has a graceful profile in<br />

addition to the excellent economic and<br />

technical viability.<br />

References<br />

[1] Code for Construction on Bridge and Calvert<br />

of Railway TB10203-2002, China Railway<br />

Publishing House, Clause 9.5.7, 87.<br />

[2] Luo S, Yan A, Liu Z. The research of long<br />

span continuous rigid frame-flexible arch hybrid<br />

bridge structure. Journal of Railway Science and<br />

Engineering 2004; 2: 57–62.<br />

Fig. 6: The completed Yichang Yangtze River Railway Bridge<br />

<strong>SEI</strong> Data Block<br />

Conclusion<br />

The PC rigid-frame, concrete-filled,<br />

steel tube arch hybrid structure<br />

adopted for the Yichang Yangtze River<br />

Bridge is a new application for construction<br />

of a double-tracked railway<br />

bridge. The PC rigid frame girder and<br />

the concrete-filled steel tube arch bear<br />

the load together; the dead load of<br />

the girder is mainly borne by the rigid<br />

frame itself; the secondary load and<br />

live load are borne jointly by the girder<br />

and the arch. The magnitude of the<br />

forces resisted by the girder and arch<br />

depends, respectively, on the rigidity of<br />

its individual structure and the area of<br />

the flexible suspenders. 2 The mechanical<br />

behaviour of the structure to resist<br />

bending is made up of resistance of the<br />

arch to the compressive stress and that<br />

of the girder to the tension stress. The<br />

horizontal thrust of the arch is balanced<br />

by the axial tension in the girder. Since<br />

the structure has excellent stiffness<br />

and stability, most external loads could<br />

not cause a horizontal thrust on the<br />

pier As a result, the bending moment<br />

resisted by girder is decreased; so the<br />

section dimensions of the girder could<br />

be reduced accordingly, maximizing<br />

the advantages of force resistance of<br />

Owner:<br />

Wuhan Railway Administration<br />

Bureau, Ministry of Railways, PRC<br />

Designer:<br />

China Railway Siyuan Survey and<br />

Design Group Co. Ltd., PRC<br />

Contractor:<br />

China Railway Major Bridge<br />

Engineering Group Co. Ltd., PRC<br />

Steel (t): 25 167<br />

Concrete (m 3 ): 13 672<br />

Total cost (USD million): 64,18<br />

Service date:<br />

Expected by<br />

December <strong>2010</strong><br />

450 Technical Report Structural Engineering International 4/<strong>2010</strong>


Static and Dynamic Analysis of the “Piedra Movediza”<br />

Replica Rock, Argentina<br />

María Inés Montanaro; Civil Eng., Maria Haydee Peralta; Civil Eng., Norma Ercoli; Civil Eng., Maria Laura Godoy;<br />

Civil Eng., Irene Rivas; Prof., National University of the Centre of Buenos Aires Province, Civil Engineering, Buenos Aires,<br />

Argentina. Contact: mmontana@fio.unicen.edu.ar<br />

Abstract<br />

This paper describes the activities<br />

involved in the development of the project<br />

“Piedra Movediza” Replica, located<br />

on the La Movediza hill in the city of<br />

Tandil, Argentina, in May, 2007. The<br />

nature of the project warranted multidisciplinary<br />

works involving topographers,<br />

geologists and private enterprises<br />

for the design and realisation of the<br />

covering material, construction of the<br />

anchors, hoisting and the subsequent<br />

assembly. Attention was especially<br />

given to the structural aspect, for which<br />

a typology consisting of an internal<br />

steel structure formed by a framework<br />

consisting of four grids arranged in two<br />

orthogonal planes was adopted. These<br />

grid beams transferred the load to the<br />

grid column or mast with its vertical axis<br />

coinciding with the vertical axis of support.<br />

The structure was complemented<br />

by cross, longitudinal and horizontal<br />

frames, which played a dual role: first,<br />

to copy the rock’s external geometry<br />

and serve as a mould for the outer covering,<br />

and second, to comply with the<br />

resistance function of transmitting the<br />

external loads to the internal structure.<br />

A static and dynamic analysis of the<br />

resistant structure was performed.<br />

Keywords: steel; composite; static<br />

and dynamic analysis.<br />

History and Aim<br />

Tandil has always been famous for<br />

the imposing Piedra Movediza, which<br />

used to rest at an inconceivably steep<br />

angle on one of the hilltops of the city.<br />

This rock finally smashed on the valley<br />

floor on 29 February 1912. The place<br />

where this rock once stood is one of<br />

the most visited places in Tandil.<br />

The Piedra Movediza (or “moving<br />

stone”), a large boulder, stood seemingly<br />

miraculously balanced on the<br />

brink of a chasm. To demonstrate the<br />

slight movements of the boulder, it<br />

was common practice to place bottles<br />

or some other things on its base to<br />

see them break. 1 The moving stone<br />

toppled on 29 February 1912. In May<br />

2007, a replica was set up in the same<br />

place where the original rock stood.<br />

Work Team<br />

An agreement was signed between<br />

the Tandil Government and the<br />

Universidad Nacional del Centro de la<br />

Provincia de Buenos Aires (UNCPBA)<br />

to commence the studies that would<br />

embark on the project to create a replica<br />

of the moving rock, a team was<br />

formed with teachers and a graduate<br />

of the Civil Engineering Department,<br />

coordinated by the engineer, and<br />

external participants. The engineer,<br />

and a team from the Universidad<br />

Nacional de La Plata, together with<br />

the Aeronautics Department and a<br />

geologist, also took part.<br />

Plan of Action<br />

A plan of action was established following<br />

certain given guidelines:<br />

– The replica should have the same<br />

geometrical dimensions as the original<br />

rock.<br />

– It should be fixed in the same position<br />

as the original rock.<br />

– It should be situated at the same<br />

place on Cerro La Movediza.<br />

– It should consist of a steel structure<br />

and it should be coated with a composite<br />

material capable of reproducing<br />

the colour and texture of the<br />

original rock.<br />

Activities<br />

Topographic Studies<br />

The planialtimetric topographic surveys<br />

determined the position and location<br />

of the replica. The geometry of<br />

the replica was determined from an<br />

analysis of the available information<br />

and from the survey of the geometry<br />

of the fallen rock that lies at the foot<br />

of the Cerro La Movediza. Infrared<br />

rays, laser electronic tachometres and<br />

GPS receiver systems were used to this<br />

end. 1523 points among the three existing<br />

pieces of rock were surveyed in a<br />

planialtimetric way. The study showed<br />

that the rock had an approximate volume<br />

of 91 m 3 , an approximate weight<br />

of 248 tons and an external surface of<br />

approximately 133 m 2 .<br />

Geological and Geotechnical Surveys<br />

These studies included, on one hand,<br />

the evaluation of the hill discontinuities<br />

existing in the location. It was<br />

observed that the hill had a fissure, the<br />

geological and geotechnical characteristics<br />

of which were studied, leading to<br />

the conclusion that this did not, in any<br />

way, render the task of anchoring the<br />

replica to the hill impossible.<br />

On the other hand, the characteristics<br />

of the rock base, which determined the<br />

depth of anchors and their size, were<br />

studied. It was reported that the solid<br />

rock on the hill had a granite structure<br />

and that it presented very good<br />

geological characteristics 2 ideal for<br />

anchoring the replica.<br />

Wind Tunnel Study<br />

This study allowed the assessment<br />

of the impact of the topography on<br />

the location and the geometry of the<br />

structure in the distribution of the<br />

wind pressure. The static and dynamic<br />

forces at the anchoring for each direction<br />

of the wind were also measured.<br />

The range of standard frequencies<br />

where higher levels of non-stationary<br />

aerodynamic load can be found was<br />

also determined.<br />

A 1 : 40 scale prototype model of<br />

the replica of the “Piedra Movediza”<br />

and part of the summit of Cerro La<br />

Movediza were tested in the wind tunnel<br />

of the Aeronautics Department<br />

at the Faculty of Engineering at the<br />

UNLP. The model of the hill summit<br />

was modified to allow its rotation<br />

in the tunnel test section and to<br />

study the winds in eight directions (S,<br />

SW, W, NW, N, NE, E and SE). The S<br />

directions (direction of more frequent<br />

winds) – SW (direction of winds of<br />

higher intensity) and SE – follow the<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 451


original geometry. On the basis of the<br />

observations at the location, it was<br />

estimated that the changes in the modelling<br />

of the hill were conservative, as<br />

they may induce higher wind loads in<br />

the test than those obtained in reality.<br />

Figure 1 shows an image of the tested<br />

model and the equipment used.<br />

For each wind direction, forces and<br />

distribution of pressure at different<br />

speeds were measured to verify the<br />

independence of the Reynolds number<br />

of dimensionless coefficients of force<br />

and pressure.<br />

The distribution of pressure in the replica<br />

model for the eight wind directions<br />

was measured in 56 nodes by means of<br />

sensors, presenting the information as<br />

a dimensionless coefficient of pressure<br />

C p . Analysis of the specified pressure<br />

coefficients reveals the strong influence<br />

of the irregular geometry and the<br />

topography (at location) in their distribution.<br />

The data reported by the tests<br />

allowed the corresponding calibration<br />

of the numerical models used and the<br />

comparison of results.<br />

Covering Material<br />

The covering material has a structural<br />

function, necessitating that certain<br />

requirements of strength, stiffness and<br />

durability be followed. Because of the<br />

nature of the project, the covering<br />

material should simulate the texture<br />

and colour of the original rock and<br />

achieve an optimum weight to facilitate<br />

the replica hoisting.<br />

From the resistance point of view, the<br />

replica should resist the pressure of<br />

the wind and transmit it to its internal<br />

resistant structure.<br />

In response to the request for simulating<br />

the texture and colour of the<br />

original rock, a large amount of previous<br />

works carried out with composite<br />

materials in this regard were referred<br />

to. Additionally, from a construction<br />

and assembly point of view, these kinds<br />

of materials have advantages because<br />

of their lower weight, and based on<br />

this, a composite material for the covering<br />

was adopted.<br />

Structural Project<br />

The project was designed considering<br />

that the replica would have the same<br />

geometrical dimensions, same appearance,<br />

texture and colour. It would also<br />

be located in the same position and<br />

location on the hill with respect to the<br />

original rock. Data from the previously<br />

indicated stages of the study made the<br />

structural project possible.<br />

Structural Typology<br />

The typology adopted, as shown in<br />

Fig. 2, consists of an internal steel structure<br />

built in a construction company<br />

and made of a framework consisting of<br />

four grids arranged in two orthogonal<br />

planes according to Fig. 3. These grids<br />

relieve in a column, with its vertical<br />

axis coinciding with that of the original<br />

rock. The structure is complemented<br />

by vertical and horizontal frames, made<br />

of stiffened steel plates, which play a<br />

dual role: first, copying the external<br />

geometry of the rock and serving as a<br />

mould for the outer covering, and second,<br />

transmitting the loads caused by<br />

the wind to the internal structure. The<br />

connection of the horizontal external<br />

frames with the grids is mainly through<br />

the beams at three planes of stiffness.<br />

Structural Analysis<br />

A software based on Finite Element<br />

Method, was used for the static and<br />

dynamic structural analysis.<br />

Fig. 2: Structural typology<br />

Static Behaviour<br />

In the first instance and for the purpose<br />

of making a pre-dimensional study 3 of<br />

the structure elements, simplified analyses<br />

in orthogonal planes coincident<br />

with those of the main grids of the<br />

frame were performed. These analyses,<br />

apart from enabling the initial predimensional<br />

study, allowed making<br />

the typologies of the framework grids<br />

consistent with the external shape and<br />

dimensions of the original rock.<br />

Simplified load combination related to<br />

adopted models, and load hypotheses<br />

arising from the overlapping of the<br />

dead load (weight structure + weight<br />

of the covering) and wind were considered.<br />

To this effect, a distributed<br />

covering weight of 3 tons and a wind<br />

pressure corresponding to a speed of<br />

50 m/s equivalent to 180 km/h was<br />

considered. The distribution of the<br />

wind pressures was obtained from<br />

the coefficients of pressure obtained<br />

through the tests in the wind tunnel for<br />

the eight tested directions and for the<br />

highest speed corresponding to 50 m/s.<br />

For the transverse grids, the overload<br />

corresponding to the north wind as the<br />

worst combination was considered.<br />

Following the pre-dimensional analysis<br />

of plane models, spatial static analyses<br />

were conducted considering nine load<br />

hypothesis corresponding to the dead<br />

loads and the combination of the eight<br />

states of independent wind loads with<br />

the dead load. The worst results for the<br />

central column corresponded with the<br />

column base plate.<br />

Dynamic Behaviour<br />

The unique characteristics of the structure<br />

were obtained by dynamic analysis<br />

for purposes of comparison with inputs<br />

from wind tunnel tests, with reference<br />

to the standard frequency range in<br />

which the highest level of energy from<br />

non-stationary aerodynamic loads is<br />

concentrated. The reported range of<br />

normalised frequencies corresponds<br />

to a 1,7 and 6,5 Hz dynamic load<br />

Fig. 1: Prototype model in wind tunnel<br />

Fig. 3: Frameworks<br />

452 Technical Report Structural Engineering International 4/<strong>2010</strong>


frequency. The first dynamic analysis<br />

performed with the pre-dimensional<br />

study used in the static analyses yielded<br />

values of the fundamental frequency,<br />

which was in the range of the highest<br />

load energy measured in the wind tunnel.<br />

This contributed to the stiffness of<br />

the column of the structure and other<br />

areas of importance that led to a fundamental<br />

frequency of 8 Hz away from<br />

the range of excitement mentioned.<br />

Design of the Foundation<br />

The design of the foundation of the<br />

moving rock replica basically includes<br />

four structural anchors of 5 m length,<br />

embedded in the base rock held by a<br />

single anchor plate of 70 mm thickness,<br />

with stiffeners arranged orthogonally<br />

and inserted into the rock with a<br />

grout-type injection material between<br />

the rock and the plate to ensure proper<br />

adherence. Additionally, three vertical<br />

inserts 1,5 m deep and of construction<br />

nature were used.<br />

To facilitate the union of the structure<br />

with the foundation, the central column<br />

was welded to a 1200 × 1150 ×<br />

50 mm base plate. This base plate was<br />

bolted, after the hoisting, to an ISO 8,8<br />

anchor plate with 20 bolts of 1 ½" ISO<br />

8,8 displayed on the anchor plate for<br />

that purpose. The base plate had corresponding<br />

holes that were perfectly<br />

aligned for the assembly.<br />

Construction Process<br />

The construction process was led by<br />

the Secretary of the Ministry of Public<br />

Works and Services for the city of<br />

Tandil. The structure was built in a<br />

construction company in Tandil.<br />

Figures 4 and 5 show the column, the<br />

grids and part of the frames, and the<br />

arrangement of the cross frames and<br />

the base plate, respectively.<br />

The outer covering was built in as<br />

specified, with some adjustments in the<br />

colour and texture.<br />

The drilling for the anchoring was<br />

done under the supervision of a geologist,<br />

taking appropriate precautions<br />

considering the particular site where<br />

the task had to be carried out. The final<br />

re-design for the arrangement of the<br />

anchors and the corresponding plate<br />

proved to be an arduous task. To this<br />

effect, a pattern of the dimensions of the<br />

anchor plate, which facilitated the rearrangement<br />

in response to the requirements<br />

of the geological survey with<br />

regard to the separation of the fissure<br />

Fig. 4: Central column<br />

Fig. 5: Framework construction<br />

plane, was built. The anchors were built<br />

by a company that specialised in rock<br />

anchoring; they also set the inserts and<br />

injected the grout according to the<br />

specifications stated. The anchor plate<br />

was hoisted by a crane already installed<br />

at the foot of the hill for mounting of<br />

the replica. The anchor plate with its<br />

20 bolts was fixed to the replica via the<br />

base plate with holes. Before the hoisting<br />

of the replica, a test of uprooting<br />

the anchors was done, with results that<br />

were within the established limits. The<br />

relocation of the replica, from the construction<br />

company to the foot of the<br />

hill, was accompanied by applauding<br />

and waving of flags by deeply touched<br />

people, which showed what the rock<br />

meant to them and to the city.<br />

The construction process lasted for<br />

approximately 4 months. The replica<br />

successfully crowned the hill on 13<br />

May 2007.<br />

A huge crane of about 9 tons hoisted<br />

the replica (Figs. 6 and 7).<br />

Conclusion<br />

The notable geometry of the replica<br />

and the site characteristics of the location<br />

made this project a real challenge.<br />

The development of the project and<br />

its subsequent execution showed the<br />

importance of working together for a<br />

project, especially when different disciplines<br />

are involved. It also made possible<br />

the sharing of knowledge between<br />

Fig. 6: The relocation of the replica<br />

Fig. 7: The replica crowned on the hill<br />

university and community, making the<br />

society highly appreciative of the university’s<br />

social role.<br />

References<br />

[1] “La Piedra Viva” Elías El Hage, Pomy Levy.<br />

Alfredo Bossio. Artes Gráficas. 1° Ed. Mayo,<br />

2007.<br />

[2] Informe petrográfico geofísico de los estudios<br />

realizados sobre la Piedra Movediza de Tandil.<br />

Inédito. Secretaría de Minería. Buenos Aires,<br />

1962.<br />

[3] Reglamento Argentino de Estructuras de<br />

Acero para Edificios. CIRSOC 301, Buenos<br />

Aires, Julio, 2005.<br />

<strong>SEI</strong> Data Block<br />

Owner:<br />

Municipalidad de Tandil (Argentina)<br />

Contractor:<br />

Metalúrgica Marcelo Fernandez<br />

Designer:<br />

National University of the Centre of<br />

Buenos Aires Province<br />

Steel (t): 7<br />

Composite (t): 3<br />

Estimated cost (EUR million): 0,2<br />

Service Date: May 2007<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 453


Footbridge Studenci over the Drava River in Maribor,<br />

Slovenia<br />

Viktor Markelj, Structural Eng., Manager, PONTING d.o.o. Maribor; Lecturer, Faculty of Civil Engineering, University of Maribor,<br />

Slovenia. Contact: viktor.markelj@ponting.si<br />

Abstract<br />

The paper presents the design and<br />

construction of the new footbridge<br />

Studenci over the Drava River in<br />

Maribor. In 2004, the City of Maribor<br />

issued an open, anonymous national<br />

call for proposals for the reconstruction<br />

of the old bridge. In spite of the<br />

fact that it was an existing bridge<br />

that was to be reconstructed, the new<br />

bridge appeared in a new, entirely different,<br />

contemporary image.<br />

The existing old bridge structure of<br />

two steel I-shaped girders with concrete<br />

deck was replaced by a new,<br />

steel, transparent, space truss structure,<br />

with a wooden deck and linear<br />

LED illumination. Only the supports<br />

from the old bridge were preserved<br />

and these also were reconstructed and<br />

strengthened.<br />

The ‘structural solution’ as presented<br />

in this paper, received a Footbridge<br />

Award in 2008.<br />

Keywords: footbridge; pedestrian<br />

bridge; reconstruction; design competition;<br />

steel structure.<br />

Introduction<br />

The new footbridge over the Drava<br />

River in Maribor is in fact a reconstruction<br />

of the old bridge where the<br />

existing supports were strengthened<br />

and the old superstructure was substituted<br />

with a new one.<br />

In spite of a relatively simple and clear<br />

structural system, the bridge is distinctly<br />

recognizable due to the interaction<br />

of its walking surface and structure<br />

elevation. Because of the transparency<br />

of its truss structure and symmetry, it<br />

is neutral to both the environment and<br />

the river landscape. In addition to the<br />

attractive configuration, the original<br />

design made way for an attractive and<br />

economic method of construction.<br />

History<br />

The history of the old footbridge is full<br />

of character. It was constructed in the<br />

year 1885 with three simply supported<br />

truss girders and wooden supports in<br />

the riverbed. After the flood in 1903,<br />

the bridge was provided with intermediate<br />

stone supports (Fig. 1). During<br />

World War II, it was destroyed twice<br />

and each time reconstructed by the<br />

army. In the year 1946, the bridge was<br />

swept away by floods; two years later,<br />

the superstructure was substituted by a<br />

new one—with two welded plated girders<br />

(Fig. 2). After the completion of the<br />

hydro power plant in 1968, the water<br />

level increased by about 5 m. In 2007,<br />

a new superstructure was constructed<br />

according to the winning solution of<br />

the design competition conducted in<br />

2004.<br />

Design of the New Footbridge<br />

To minimize construction costs, it was<br />

necessary to reuse the existing supports<br />

of the old footbridge, thus dictating<br />

the spans of 3 × 42 = 126 m,<br />

and, at the same time, it was necessary<br />

to increase the existing navigation<br />

clearance under the bridge from 3,0 to<br />

3,6 m of clear opening. Beside common<br />

requirements such as safety, durability<br />

and economy, the city of Maribor also<br />

sought an attractive, unique bridge<br />

that would be in harmony with the<br />

environment and would prove appealing<br />

to the town dwellers.<br />

The selected solution was a triangular<br />

steel truss as a primary longitudinal<br />

structure (spine) along which a<br />

secondary transversal structure was<br />

to be raised (ribs). As neither of the<br />

banks had sufficient height for the<br />

structure, the structure was raised<br />

upwards through the walking surface,<br />

thus being divided into two parts.<br />

With a gradual rise of the deck, the<br />

structure disappears under the floor,<br />

joining both walkways over the middle<br />

of the river and providing a common<br />

surface for an unimpeded view<br />

of all sides and a peaceful meeting<br />

place.<br />

Fig. 1: Footbridge Studenci from 1885 to 1946 Fig. 2: Footbridge Studenci from 1948 to 2007<br />

454 Technical Report Structural Engineering International 4/<strong>2010</strong>


Longitudinal section / side view<br />

1<br />

5,25<br />

42,00<br />

2<br />

136,50<br />

42,00<br />

3<br />

42,00<br />

4<br />

5,25<br />

Studenci<br />

Taborsko nabrezje<br />

Koblarjev zaliv<br />

Lent<br />

Ruska cesta<br />

Strma ulica<br />

3,05<br />

3,60<br />

Drava river<br />

3,05<br />

Support axis 1<br />

Support axis 2<br />

Plan / view<br />

Support axis 3<br />

Support axis 4<br />

5,25<br />

42,00<br />

136,50<br />

42,00 42,00<br />

5,25<br />

3,20<br />

5,80<br />

Drava river<br />

Fig. 3: Longitudinal section and plan of the new footbridge (Units: m)<br />

Description of the Footbridge<br />

and the Structure<br />

The footbridge axis is in a straight<br />

line, the longitudinal alignment is<br />

in a convex vertical rounding with<br />

R = 1045 m and with a maximal slope<br />

of 5%. In the middle of the footbridge,<br />

the deck width is 3,20 m and at<br />

either end it is split into two parts of<br />

2,40 m each, divided by a visible main<br />

structure breaking through the deck<br />

(Fig. 3).<br />

The variation of the structure and deck<br />

alignment was solved by dividing the<br />

structure into two systems supporting<br />

the deck:<br />

– The spine—main steel structure is a<br />

centrally positioned triangular truss<br />

of a constant structural height and<br />

width. The geometry enabled a simple<br />

production in a workshop: after<br />

assembling on site by welding, it was<br />

erected in place simply by incrementally<br />

launching it over the existing<br />

structure. The main steel structure<br />

is made of structural steel S355 and<br />

protected with anticorrosive coats in<br />

light grey.<br />

– Ribs—secondary steel structure rises<br />

along the main structure and carries<br />

the walking surface. Erected<br />

by bolting after the erection of the<br />

main structure, it is protected by hot<br />

galvanizing.<br />

– Wooden deck structure made of<br />

transversal boards from hard exotic<br />

wood of 42 mm thickness, with a<br />

sawtooth profile on the top.<br />

The main structure is a space steel<br />

truss, consisting of three longitudinal<br />

Secondary structure<br />

cross bracket<br />

1,760<br />

Cross bracket width<br />

f298,5 mm<br />

d = 8–2 mm<br />

B = 3,200–5,800<br />

Bridge axis<br />

Main structure attachment<br />

plates (altering position)<br />

f298,5 mm<br />

d = 26–20 mm<br />

f114,3 mm<br />

d = 8–20 mm<br />

1,500<br />

0,720–3,140<br />

Main structure width<br />

1,750<br />

1,760<br />

Cross bracket width<br />

Fig. 4: Cross section of the footbridge is of constant shape—changing only at the ends<br />

of the secondary structure<br />

New pier cap<br />

Micropile<br />

f25 cm, L = 15,0 m<br />

Strengthened existing pier<br />

min. 7,000 m<br />

Jet-grouting piles<br />

0,25<br />

1,60<br />

Pier axis<br />

2,70<br />

3,50°<br />

1,57<br />

0,25<br />

1,75° 1,75°<br />

254,524<br />

New bearing block<br />

0,50<br />

253,200<br />

14,50<br />

15,00<br />

Water level<br />

River bed<br />

New pier cap<br />

Micropile<br />

f25 cm, L = 15,0 m<br />

Strengthened existing pier<br />

Fig 5: Reconstruction of the old river piers with jet-grouting piles (Units: m)<br />

min. 7,000 m<br />

Jet-grouting piles<br />

0,25<br />

7,50°<br />

os brvi<br />

5,45<br />

6,50<br />

0,25<br />

7,50° 7,50°<br />

3,17 3,17<br />

0,25<br />

7,50°<br />

New bearing block<br />

254,524<br />

0,50<br />

253,200<br />

14,50<br />

15,00<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 455


30,0°<br />

30,0°<br />

Original (glued) wooden section 240/80 mm<br />

240,0<br />

30,0°<br />

30,0°<br />

LED linear diodes (L = 2400 mm, e = 40 mm)<br />

pipes—one pipe for the upper chord<br />

and two for the bottom chord. The<br />

triangular cross section is of constant<br />

form; the change is only in the<br />

pipe thickness and the position of the<br />

accessory piece for bolting the secondary<br />

structure (Fig. 4). The axial<br />

distance between the upper and bottom<br />

pipes is 1,75 m which gives a total<br />

structural height of 2,05 m; the axial<br />

distance between the lower flanges is<br />

1,50 m. The longitudinal pipes are of<br />

diameter ϕ = 298,5 mm, wall thickness<br />

varies from 8 to 20 mm and for<br />

the diagonal and cross pipes, ϕ = 114,3<br />

mm. The camber of the main structure<br />

is of a constant radius R = 4000 m,<br />

the upright connections between layers<br />

are not vertical but radial, so that<br />

the element lengths and mutual angles<br />

are equal along the total bridge length,<br />

making the construction cheaper and<br />

simpler.<br />

The old river piers were reinforced<br />

with six piles and the abutments—<br />

because of greater width—with nine<br />

piles. The total length of the jet-grouting<br />

piles is 15 m—8 m through the old<br />

pier structure and 7 m through the<br />

gravel base (Fig. 5). The tops of existing<br />

piers have been adapted for the<br />

needs of the new structure.<br />

80,0<br />

Polycarbonate protected<br />

strip d = 3 mm<br />

Fig 6: Footbridge lighting is concealed within the wooden top rail of the steel railing<br />

Fig. 7: Main truss of new footbridge positioned over the old structure<br />

Fig. 8: Dismantling of the old structure<br />

456 Technical Report Structural Engineering International 4/<strong>2010</strong>


method of erection was possible. After<br />

erection, the new structure was supported<br />

by the remainder of the existing<br />

structure. Truss elements were erected<br />

above the intermediate supports and<br />

the bearings were grouted. Then the<br />

old footbridge was dismantled (Fig. 8).<br />

Some details of the new footbridge<br />

are shown in Fig. 9. The profile of the<br />

reconstructed footbridge Studenci can<br />

be seen in Fig. 10.<br />

Fig. 9: Some details of the new footbridge<br />

Conclusion<br />

Since the city of Maribor could not<br />

give up more than 120 years old<br />

pedestrian route, it decided to reconstruct<br />

the old destroyed footbridge -<br />

on the existing supports, but with<br />

the new superstructure. The original<br />

design of the bridge’s reconstruction<br />

provides an attractive form and at<br />

the same time enables a very costeffective<br />

way of building. The design<br />

solution has been presented with the<br />

Footbridge Award 2008 in the category<br />

of Technical medium span, with<br />

the judges declaring it as »A pleasing<br />

technical solution to a difficult set of<br />

criteria« and noticing its »Interesting<br />

new technical ideas«.<br />

<strong>SEI</strong> Data Block<br />

Fig. 10: Reconstructed footbridge Studenci over the Drava River in Maribor<br />

The secondary structure, supporting<br />

the wooden deck structure, is<br />

composed of transverse cantilever<br />

girders and longitudinal girders and<br />

is screwed to the main structure all<br />

along. The transverse cantilevers<br />

are at the same time railing balusters<br />

and are equal along the entire<br />

bridge length. The deck is made of<br />

transversely placed bangkirai wooden<br />

boards. The discreet lighting from the<br />

handrail is another special feature of<br />

the footbridge. The installed lighting<br />

power of LED diodes is only 350 W<br />

(Fig. 6).<br />

Construction<br />

The footbridge construction was innovative<br />

and was built at a low cost, as it<br />

used the existing structure as a support<br />

for erection. The structure was divided<br />

into two parts, the main longitudinal<br />

one and the secondary transverse one.<br />

The main structure is a triangular truss<br />

that was welded by segments and progressively<br />

positioned over the existing<br />

structure (Fig. 7). The separation<br />

into main and secondary structures<br />

reduced the transverse size of the<br />

structure to such an extent that this<br />

Owner:<br />

City of MARIBOR, Slovenia<br />

Structural design:<br />

PONTING d.o.o., Mari bor, Slovenia<br />

Contractors:<br />

SGP Pomgrad GNG d.o.o., Slovenia<br />

Steel structure: Meteorit d.o.o., Slovenia<br />

Bridge type: Steel space truss<br />

Bridge size:<br />

Length: 130 m,<br />

Width:<br />

3,2–5,8 m<br />

Deck:<br />

Bangkirai wood: A = 550 m 2<br />

Steel S355:<br />

93 000 kg<br />

Lighting:<br />

LED 350W<br />

Total cost (EUR million): 1,2<br />

Service date: December 2007<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 457


Sanchaji Bridge: Three-Span Self-Anchored Suspension<br />

Bridge, China<br />

Gonglian Dai, Prof., Dr, Civil Eng.; Xuming Song, Lecturer, Dr, Civil Eng.; Nan Hu, Graduate Student; Central South University,<br />

Changsha, China. Contact: daigong@vip.sina.com<br />

Abstract<br />

Sanchaji Bridge across the Xiangjiang<br />

River is located at the northern section<br />

of the Second Ring Road Project<br />

in Changsha, Hunan Province, China.<br />

Its main span is the longest in the<br />

world among all self-anchored suspension<br />

bridges constructed using double<br />

towers and double cables till now.<br />

For a self-anchored bridge, structural<br />

behaviour and construction methods<br />

are totally different from those of a<br />

traditional suspension bridge. A main<br />

cable containing 37 prefabricated<br />

strands and streamlined steel box cross<br />

section was used as stiffened girder<br />

with a height of 3,6 m. A fully welded<br />

anchoring chamber was adopted to<br />

connect the main cable with the stiffened<br />

girder for the first time. The main<br />

tower is of reinforced concrete with a<br />

variable hollow box section. As for the<br />

construction method, launching methods<br />

were selected for the erection of<br />

steel box girder and “non-stress”<br />

method for installation of hangers. The<br />

construction of the bridge was started<br />

on 10 September 2004 and completed<br />

on 1 September 2006. The completion<br />

of the bridge effectively relieved the<br />

traffic pressure in the northern region<br />

of Changsha and played a vital role in<br />

improving local economy. This paper<br />

introduces several features of the<br />

Sanchaji Bridge, including type selection,<br />

structural behaviour and construction<br />

methods.<br />

Keywords: self-anchored suspension<br />

bridge; structural system; design<br />

parameters; construction method.<br />

a (5 × 65 m) PSC continuous beam.<br />

Before project bidding, there were<br />

five bridges across Xiangjiang River in<br />

Changsha, including two arch bridges,<br />

two beam bridges and a cable-stayed<br />

bridge. The self-anchored suspension<br />

bridge system was selected because<br />

of its logical structural principles and<br />

the aesthetic value addition to the<br />

landscape 1 .<br />

Geological Condition<br />

The Sanchaji Bridge is located on a quaternary<br />

stratum, which mainly contains<br />

soil in the upper and slate in the lower<br />

levels. There are not many ups and<br />

downs at the surface of the river bed,<br />

where the main ingredients are sandy<br />

slate and metamorphic sandstones.<br />

The uniaxial compressive strength is<br />

20,9 to 65,60 MPa. Rock forms the<br />

fourth level. The arrangement of the<br />

main span and the geological conditions<br />

are shown in Fig. 2. It can be seen<br />

that piers 11 and 12 that formed the<br />

foundation for the two towers adopted<br />

drilled pile groups of C30 class concrete<br />

with length between 16 and 24 m.<br />

The C30 concrete pile cap under each<br />

tower is a dumbbell-shaped structure<br />

including two caps with diameter of<br />

17 m and thickness of 5 m linked by<br />

7 × 4 m tie beam. Under each tower,<br />

there are 18 piles with a diameter of<br />

2,4 m. The width of the channel located<br />

in the west of the island is about 700<br />

m and its greatest depth is at about<br />

6,0 m. Flood peak of Xiangjiang River<br />

frequently appears from late April to<br />

June, which covers about 86% of total<br />

number of peaks occurring in 1 year.<br />

The basin is so large that the flooding<br />

is quite heavy.<br />

Design Criteria<br />

The main design criteria for Sanchaji<br />

Bridge are high-way-I with design<br />

speed of 60 km/h; three-lane traffic<br />

loading in each direction with pedestrian<br />

load of 4 kN/m; deck width of 35 m;<br />

lateral slope of bridge deck is 2,0% in<br />

two directions; longitudinal slope of<br />

bridge is 1,5%; temperature: design<br />

normal temperature is 20°C with mutative<br />

range of 0 to 40°C; normal design<br />

wind speed refers to average speed of<br />

25,9 m/s; designed to withstand upto 7<br />

grade intensity of magnitude for earthquake;<br />

anti-collision standard refers<br />

to third-level fairways with 400 kN<br />

Introduction<br />

Sanchaji Bridge shown in Fig. 1 is<br />

located at downstream Xiangjiang<br />

River in the northern region of<br />

Changsha, where the width between<br />

river banks is about 1442 m. The total<br />

length of the main bridge is 1577 m,<br />

with span arrangement of the whole<br />

bridge containing a (8 × 65 m) prestressed<br />

concrete (PSC) continuous<br />

beam, a (70 + 132 + 328 + 132 + 70 m)<br />

self-anchored suspension bridge and Fig. 1: The close view main span of Sanchaji Bridge<br />

458 Technical Report Structural Engineering International 4/<strong>2010</strong>


K41 + 504,35 K42 + 236,35<br />

70 12 + 12 × 9 + 12 = 132<br />

732<br />

11 + 34 × 9 + 11 = 328<br />

12 + 12 × 9 + 12 = 132 70<br />

129,655<br />

129,655<br />

with a spacing of 600 mm. The thickness<br />

of the inclined web flange is 8 mm<br />

with longitudinal closed rib stiffeners<br />

at 400 mm spacing.<br />

55,770<br />

56,720<br />

57,964<br />

58,515<br />

57,964<br />

56,720<br />

55,770<br />

Main Tower and Pier<br />

23,496<br />

9 10 11 12 13 14<br />

Fig. 2: The main span arrangement of Sanchaji Bridge (Units: m)<br />

in longitudinal and 550 kN in lateral<br />

direction; channel clearance is 10 m<br />

while flood level is 36,78 m.<br />

Structural System<br />

Since no bridge of a similar type was<br />

constructed ever before, 1 : 28 overall<br />

model test and 1 : 5 steel anchoring<br />

chamber test were carried out to<br />

verify the accuracy of finite element<br />

analyses. Test results were used in conceptual<br />

design and the details of the<br />

design of the bridge are as follows:<br />

(a) a five-span continuous girder was<br />

used in a self-anchored suspension<br />

bridge for the first time, which eliminated<br />

tremendous uplift force created<br />

by the anchorage of main cable at #10<br />

and #13 piers. In this way, a negative<br />

reaction force at the bearings could<br />

be basically avoided under the dead<br />

load, and very little rotation at the<br />

beam end occurred during the service<br />

period; (b) the issue of stability is<br />

vital in this bridge because the cable is<br />

anchored at the girder of the suspension<br />

bridge. So, it was addressed by<br />

placing precast concrete blocks in the<br />

box girder near the section of the pier<br />

where the heaviest weight was 8000 kg<br />

at #10 and #13 piers and 6000 kg at #11<br />

and #12 piers. In this way, there was no<br />

negative reaction force at the bearings<br />

of these four piers mentioned under<br />

both dead load and live load, and<br />

the safety coefficient was maintained<br />

above 1,3; (c) the selection of restraint<br />

type would have a huge impact on<br />

structural behaviour. On the basis of<br />

analysis and comparison, four MSTU<br />

25,202<br />

dampers were installed between the<br />

stiffened girder and tower columns<br />

at both towers, where each damper<br />

required a force of 100 tons. Therefore,<br />

the stiffened girder would be a floating<br />

system along the bridge under<br />

some slow loads, such as temperature<br />

variation, while the girder would be<br />

controlled by dampers at columns of<br />

the tower and reduce the force on the<br />

foundation of the tower during special<br />

live loads, such as vehicle braking<br />

force and earthquakes.<br />

Stiffened Girder<br />

In a self-anchored system, the stiffened<br />

girder will be under a great axial force.<br />

Therefore, the cross section of girder<br />

adopted a streamlined steel box type<br />

with Q345d steel, after the wind tunnel<br />

test, as shown in Fig. 3. The main<br />

dimensions of the steel box girder<br />

cross section were as follows: the clear<br />

height of the box girder at the bridge<br />

axis was 3,60 m and overall width was<br />

35 m. So, the high-span ratio of the box<br />

girder is 1 : 91,1 and high-width ratio<br />

is 1 : 9,72. The bridge deck adopted<br />

an orthotropic plate with a thickness<br />

of 12 to 14 mm at the top plate, 12 to<br />

16 mm thickness at the web plate and<br />

10 mm at the bottom plate, except<br />

at some local section near the tower.<br />

There is a diaphragm every 3 m with<br />

a thickness of 10 mm at normal section<br />

and 16 mm at the position of<br />

bearing. The U-shaped closed rib deck<br />

stiffeners are 260 mm deep and 8 mm<br />

thick with a spacing of 600 mm. At<br />

the bottom flange, the dimensions are<br />

190 mm (depth) and 6 mm (thickness)<br />

The main tower adopted a C50 class<br />

RC structure with variable cross sections,<br />

as shown in Fig. 4. The height<br />

of the tower is 106,16 m in #11 and<br />

104,453 m in #12 piers (from the top<br />

of pile cap). Above the bridge deck,<br />

the height is about 71,752 m. The cross<br />

section of the tower is of a hollow box<br />

type with dimensions of 5 (along the<br />

bridge) × 3,5 m (along the section) at<br />

the top. The column width along the<br />

direction of the bridge expands from<br />

the top to bottom according to the<br />

ratio 1 : 100. The two columns of the<br />

tower are linked by crossbeams with a<br />

spacing of 25 m at the top. There are<br />

two PSC crossbeams between columns<br />

with a hollow rectangular cross section.<br />

The crossbeam below the deck<br />

is 4,0 × 6,0 m in cross section and 0,6<br />

m thick, equipped with 48,12-ϕj15,24<br />

strands. As for the upper crossbeam,<br />

the dimensions are 3,0 × 4,5 m, with<br />

a thickness of 0,5 m, equipped with<br />

eight 12-ϕj15,24 strands. The thickness<br />

of the column is variable, being<br />

700 mm above the lower crossbeams<br />

and 1 m below the same. Furthermore,<br />

all the joint parts between crossbeams<br />

and columns are strengthened, with a<br />

gradual variation in thickness. Because<br />

of saddle installation, the solid column<br />

area was adopted with a length of 4 m.<br />

1000 2900 2200<br />

35 000/2<br />

11 400<br />

2070 1200<br />

3600<br />

3693<br />

13 807<br />

Fig. 3: The semi cross section of the bridge after the wind tunnel test (Units: mm)<br />

Fig. 4: The main tower above the bridge<br />

deck<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 459


Lightning protection devices were set<br />

up at the top of the column. Except<br />

for the two main towers, the remaining<br />

piers for approach are of twin-column<br />

type with a diameter of 3 m.<br />

Cable System<br />

On account of the larger rise–span<br />

ratio than that of the gravity-anchored<br />

suspension bridge or single tower one,<br />

it is vital to take into consideration<br />

the selection of hanger clamps, shape<br />

of cable design, installation method<br />

of hanger, and so on. The rise–span<br />

ratio of Sanchaji Bridge is 1/5 at midspan<br />

and 1/10,6245 at side span 2 . 37<br />

parallel wire strands, comprised of<br />

127 j 5,1 mm high-strength galvanised<br />

steel wires, formed one main cable,<br />

which was installed by Prefabricated<br />

Parallel Wire Strands (PPWS) method.<br />

The total length of the main cable was<br />

680,70 m with a spacing of 25,0 m<br />

between the wires. The amount of steel<br />

in the main cable was about 1026 t. The<br />

shape of the main cable was catenary<br />

at every construction stage before<br />

operation, because the shape between<br />

two tower saddles was just caused by<br />

the dead load along the cable. The<br />

material adopted for strand anchor<br />

tube and cover plate was ZG310-570<br />

and Q235-A, while the filled material<br />

in tube was zinc–copper alloy.<br />

The fully welded anchoring chamber<br />

was adopted for connecting the main<br />

cable with the stiffened girder because<br />

the height of girder was not required<br />

to increase in order to ensure superior<br />

appearance and convenient construction<br />

of the girder. Moreover, the<br />

structural behaviour of this anchor-<br />

ing chamber is superior to the concrete<br />

one because cracking caused by<br />

huge local compressive stress could<br />

be avoided. Two hanger ropes were<br />

installed at each cable band, as shown<br />

in Fig. 5. Except for the two pairs of<br />

rigid hangers near the #10 and #13<br />

piers, all hanger ropes were made of<br />

85j 5,1 mm galvanised steel wire. The<br />

prefabricated parallel wire strand<br />

hanger was pinned by a socket with<br />

cable and anchored at the girder. There<br />

are 122 pairs of hangers in the whole<br />

bridge with a spacing of 9 m. Screws<br />

with bolts were used for the connections<br />

between the hanger and girder,<br />

so that the flow of force would be<br />

clear and installation easy. The hanger<br />

clamp is a symmetrical steel casting<br />

component and the length of the clamp<br />

is controlled by declining cable force<br />

and clamp traction force. On the basis<br />

of an experiment on hanger clamp,<br />

the traction force is maintained at 12<br />

MPa instead of the regular 10 MPa to<br />

avoid the huge length of clamp. There<br />

are four saddles in the whole bridge<br />

with the self-balancing system so that<br />

blocks at the top of the tower are not<br />

needed. The major material of saddle<br />

is cast GZ25 with a weight of 40,5 t,<br />

including the upper component, lower<br />

plate, filled panel and so on.<br />

Construction Method<br />

The whole construction period was<br />

divided into five stages: foundation<br />

and tower erection; incremental<br />

launching of stiffened girder; erection<br />

of main cable; installation of hang-<br />

ers; deck system pavement. The major<br />

difficulties during the construction<br />

included four aspects as follows: (a)<br />

Incremental launching of the steel box<br />

girder. The launching platform was<br />

set between #9 and #10 piers where<br />

the water was deep and four temporary<br />

piers were placed for carrying<br />

the main girder at middle span. The<br />

standard prefabricated segments were<br />

divided into two types (12 and 9 m)<br />

and assembled in site by welding. The<br />

weight of standard segment was 195,8 t<br />

of 12 m and 152,6 t of 9 m. According<br />

to the requirement of waterway during<br />

the construction, temporary piers<br />

were placed with a maximum span of<br />

77 m, and the length of launching nose<br />

was 48 m. (b) The line shape control<br />

of the main girder and cable. Since the<br />

main girder was on a vertical curve<br />

with 24402,745 m radius, the segments<br />

were installed on the platform according<br />

to the line shape of the bridge<br />

girder and pre-camber consideration.<br />

Owing to a larger rise–span ratio,<br />

there was a huge difference in line<br />

shape before and after the installation<br />

of hangers. So, the installation of main<br />

cables should follow strict limits and<br />

enable the controlling plate to adjust<br />

the errors during cable fabrication. (c)<br />

“Non-stress” method in installation<br />

of hangers. To install the “non-stress”<br />

hangers, the girder was lifted by four<br />

temporary piers (740 mm raised in<br />

two piers near the tower and 1,48 m<br />

in piers of middle span), as shown in<br />

Fig. 6. After the installation of the<br />

hangers was completed, the temporary<br />

piers were removed and the girder<br />

system could find its final position,<br />

Fig. 5: The hanger rope between main cable<br />

and stiffened girder<br />

Fig. 6: Sanchaji Bridge after the installation of hanger (during construction)<br />

460 Technical Report Structural Engineering International 4/<strong>2010</strong>


Fig. 7: The view of Sanchaji Bridge at night in service period<br />

the transformation of the structural<br />

system was completed. In this way, the<br />

time consuming hanger tension process<br />

could be avoided and construction<br />

speed could be accelerated. It took<br />

only a week to complete the installation<br />

of 244 hangers in Sanchaji Bridge.<br />

(d) Concrete pouring and pumping.<br />

The hydration problem of huge concrete<br />

structures such as pile caps and<br />

crossbeams was managed by adjusting<br />

the mix ratio and the water circulation<br />

system inside and heat insulation<br />

system outside the concrete. Concrete<br />

pumping to high places such as the top<br />

of the main tower was addressed by<br />

improving workability of concrete by<br />

adding silica to the pipeline.<br />

Conclusion<br />

During the design process of Sanchaji<br />

Bridge, a series of theoretical analyses<br />

and model test were carried out so that<br />

it provided a solid basis for the construction.<br />

Although sound structural<br />

performance and attractive appearance<br />

make self-anchored suspension<br />

bridge a competitive proposal within<br />

a certain span, such bridge types have<br />

not come so much into practice yet.<br />

Sanchaji Bridge (Fig. 7), as the longest<br />

self-anchored suspension bridge in the<br />

world, will provide a valuable experience<br />

and reference for the design and<br />

construction of such bridge types in<br />

the future.<br />

References<br />

[1] Song X, Dai G, Fang S. Conceptual design<br />

and construction method of Sanchaji Bridge.<br />

18th China Bridge Engineering Symposium,<br />

Tianjin, China, 2008 (in Chinese).<br />

[2] Song X, Dai G. The main cable shape control<br />

and design of Sanchaji bridge. IABSE Rep. 2007;<br />

93: 586–587.<br />

<strong>SEI</strong> Data Block<br />

Owner:<br />

The Administration of 2nd Ring Road<br />

(China)<br />

Contractor:<br />

The 5th Co. Ltd of China Major Bridge<br />

Engineering Group<br />

Designer:<br />

Changsha Planning and Design<br />

Institute Co. Ltd/Central South<br />

University (China)<br />

Steel (t): 13 182<br />

Concrete (m 3 ): 18 970<br />

Estimated cost<br />

(USD million):<br />

approx.51,8<br />

Service date: September 2006<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 461


The First Extradosed Bridge in Slovenia<br />

Viktor Markelj, Structural Eng., Manager, PONTING d.o.o. Maribor; Lecturer, Faculty of Civil Engineering,<br />

University of Maribor, Slovenia. Contact: viktor.markelj@ponting.si<br />

Abstract<br />

This paper presents the conceptual<br />

design and technical solutions for the<br />

bridge over the artificial lake of Ptuj<br />

(power plant reservoir on the Drava<br />

River) on the new south main road,<br />

which connects the city of Ptuj with the<br />

Maribor–Zagreb motorway.<br />

The horizontal axis of the 430,0 m long<br />

and 18,0 m wide bridge has a radius of<br />

curvature of R = 460,0 m. The bridge is<br />

designed as an “extrados bridge,” representing<br />

an intermediate form between<br />

girder bridges and cable-stayed bridges.<br />

Structural spans of the bridge are 65 +<br />

100 + 100 + 100 + 65 = 430 m. In the cross<br />

section the bridge is a mono-cellular,<br />

box-shaped, longitudinally prestressed<br />

concrete girder; H = 2,70 m, with classic<br />

prestressing tendons inside and additional<br />

exterior extrados cables over the<br />

low pylons.<br />

The conceptual design won the open<br />

design competition in 2004 and the final<br />

tender design was completed in March<br />

2005. The construction of the innovatively<br />

designed “extrados bridge” started<br />

in November 2005, and the bridge was<br />

opened for traffic in May 2007.<br />

Keywords: extrados bridge; prestressed<br />

concrete; stay cables; deviator; saddle.<br />

Introduction<br />

In May 2004, the municipality of<br />

Ptuj and Motorway Company in the<br />

Republic of Slovenia announced an<br />

anonymous competition to select the<br />

best design solution for a bridge over<br />

the Drava River near Ptuj, the oldest<br />

city in Slovenia. On the basis of the<br />

winning design, the tender design was<br />

prepared in 2005; later in September<br />

that year, and the construction contract<br />

for EUR 8,8 million was signed.<br />

The bridge, named “Puch Bridge” after<br />

the famous Slovenian inventor, was<br />

opened to traffic in May 2007.<br />

Environmental Conditions<br />

and Other Restrictions<br />

The Ptuj Lake is the largest artificial<br />

lake in Slovenia, with a length of<br />

over 5 km, width up to 1,2 km and<br />

depth up to 15 m. At the location of<br />

the bridge, the lake is 250 m wide and<br />

about 5 m deep. The surrounding terrain<br />

is very flat. The main problem was<br />

the lake itself, particularly the shores<br />

that house all the facilities and equipment<br />

for the operation of the reservoir<br />

(safety embankments, sealing curtain<br />

and drainage), the sewer pipes on both<br />

sides and the stream beyond the right<br />

bank of the lake.<br />

For specific locations, it was necessary<br />

to take into account the relative vicinity<br />

of the old town of Ptuj with its old<br />

castle. This being a site of historical<br />

heritage, very strict restrictions were<br />

imposed in order to preserve the views<br />

of the old city. This limited the height<br />

of structures (in this case, pylons) to a<br />

maximum of 10 m.<br />

The following were other significant<br />

restrictions and conditions that<br />

affected the technical design of the<br />

bridge:<br />

– severe restrictions on the support<br />

layout due to the slurry wall on the<br />

banks of the lake;<br />

– road geometry with a sharp radius<br />

of curvature of R = 460 m;<br />

– low elevation of the bridge and the<br />

required waterway and shipping<br />

clearance of 4,0 m underneath;<br />

– difficult foundation conditions in<br />

the lake because of the rocky bed<br />

located 20 to 30 m below the water<br />

level;<br />

– other obstacles (existing sewage and<br />

other municipal water ditches and<br />

streams, the existing discharge facility,<br />

sharp crossing of the road with<br />

the left bank of the river);<br />

– two-way carriageway of width<br />

8,10 m and a separate lane of width 2<br />

× 3,10 m for pedestrians and cyclists.<br />

Design Concept<br />

The design concept of the bridge was<br />

the result of a search for the optimal<br />

response to the very difficult conditions<br />

for spanning the lake. Owing to<br />

the many restrictions, the bridge concept<br />

required large spans, but at the<br />

same time, a relatively shallow structure<br />

was necessary.<br />

The longitudinal disposition with the<br />

spans of 65 + 100 + 100 + 100 + 65 =<br />

430 m proved to be the best solution<br />

for bridging the obstacles. There was<br />

no dilemma in arriving at the cross<br />

section: taking into consideration the<br />

road curvature and torsion, bridge<br />

installation and maintenance, a box<br />

girder of maximum feasible height of<br />

2,70 m was selected. The slenderness<br />

ratio L/H = 100 m/2,7 m = 37 was obviously<br />

too large for a normal concrete<br />

continuous girder structure of constant<br />

depth. The usual prestressed concrete<br />

girder with a variable height, constructed<br />

by the free cantilever method,<br />

would require a girder height of 5,0 to<br />

5,8 m (L/20–L/17) at the support. This<br />

would close the navigation channel<br />

and was not acceptable for this reason.<br />

It was therefore necessary to support<br />

the structure from above, for example,<br />

by cables through the pylon. However,<br />

taking into consideration the cultural<br />

heritage of the region and the need to<br />

protect the view of the city, a restriction<br />

of a maximum pylon height of<br />

10 m was imposed on structural elements<br />

located on the upper side of the<br />

carriageway.<br />

The solution to this problem was the<br />

use of a new system in bridge building,<br />

the so-called “extrados” bridge, which<br />

is a kind of intermediate step between<br />

the cable-stayed and the girder bridges.<br />

In this case, the bridge girder could<br />

be more slender than a normal continuous<br />

girder, and the pylon could be<br />

lower than the pylon for cable-stayed<br />

systems.<br />

Structure Description<br />

The bridge length between the expansion<br />

joints is 433 m and the total width<br />

is 18,70 m. The static system is a continuous<br />

externally prestressed box structure<br />

2,7 m high, constructed according<br />

to the so-called system of the “extrados”<br />

bridge with structural spans of 65<br />

+ 100 + 100 + 100 + 65 = 430 m. In its<br />

entire length, the bridge lies in the curvature<br />

R = 460 m (Fig. 1).<br />

The entire structure is continuous with<br />

expansion joints only on abutments.<br />

Two central supports have longitudinally<br />

fixed bearings and carry all horizontal<br />

loadings, while the transverse<br />

horizontal loadings are carried by all the<br />

462 Technical Report Structural Engineering International 4/<strong>2010</strong>


430<br />

65<br />

100<br />

100<br />

100<br />

65<br />

1 2<br />

3<br />

4<br />

5 6<br />

Channel River Drava lake Stream<br />

1<br />

2<br />

100<br />

3<br />

430<br />

100<br />

4<br />

100<br />

5<br />

6<br />

65<br />

65<br />

3,1+01,80<br />

3,2+01,80<br />

3,0+01,80<br />

3,3+01,80<br />

2,9+36,80<br />

3,3+66,80<br />

Fig. 1: Spans 65 + 100 + 100 + 100 + 65 = 430 m in curvature R = 460 m<br />

supports. Three of the four intermediate<br />

supports are located in the reservoir.<br />

The Substructure and Foundation<br />

The substructure consists of two abutments<br />

and four intermediate piers,<br />

three of which are located in the reservoir<br />

and one on the shore. All the<br />

supports are on deep foundation with<br />

piles of diameter Φ 1,50 m. Owing to<br />

the low and relatively stiff supports, a<br />

rigid connection between the piers and<br />

the girder is not possible; however, all<br />

the supports have bearings at the top.<br />

Supports 3 and 4 have fixed pot bearings<br />

at the top, while supports 1, 2 and<br />

5, 6 have longitudinally movable and<br />

transversally fixed bearings.<br />

The bearings summary is given in<br />

Table 1.<br />

The intermediate piers are relatively<br />

low, 2,4 m thick and about 8 m wide,<br />

of a full cross section and are fixed<br />

into the pile head at the bottom<br />

(Fig. 2).<br />

At the foundation in the lake, the soil<br />

mechanics data had to be taken into<br />

account as well as the execution feasibility,<br />

structure maintenance and the<br />

hydraulic consequences. For piers, the<br />

foundation on eight piles of diameter<br />

Φ1,50 m with the oval-shaped “floating”<br />

pile head was foreseen in the<br />

design. As the superstructure lies on<br />

a curve, the piles are arranged asymmetrically<br />

so that they are uniformly<br />

loaded under the permanent load (the<br />

intermediate supports are more loaded<br />

on the inner side of the curvature—on<br />

the bearings, the ratio is 58–42%). At<br />

the pile head surface, it was simple<br />

to perform the temporary supporting<br />

during the cantilever construction.<br />

Support in axis<br />

Bearing (kN)<br />

Outside curve<br />

Inside curve<br />

Abutment1 PNe 6100 PNe 5400<br />

Pier 2 PNe 21 700 PNe 28 200<br />

Pier 3 PN 23 000 PN 28 200<br />

Pier 4 PN 23 000 PN 28 200<br />

Pier 5 PNe 21 700 PNe 28 200<br />

Abutment 6 PNe 6100 PNe 5400<br />

PNe—pot bearing movable in longitudinal direction. PN—pot bearing fixed.<br />

Table 1: Bridge bearings<br />

Superstructure<br />

The superstructure consists of three<br />

main elements:<br />

– girder roadway structure;<br />

– low pylons;<br />

– inclined cables.<br />

The roadway structure is a trapezoidal<br />

prestressed RC box of structural<br />

height 2,70 m (Fig. 3). The web thickness<br />

is 0,50 m, the upper slab is 18,16 m<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 463


3,0%<br />

Drazenci<br />

Ptuj<br />

3,0%<br />

Fig. 2: Characteristic river pier<br />

wide and the bottom slab 9,10 m. The<br />

slab thickness varies from 0,22 to 0,50<br />

m. The cantilever is 4,24 m long, at the<br />

fixing point 0,50 m thick and rebar<br />

reinforced. Owing to the flow of compression<br />

stresses, the thickness of the<br />

lower slab is enlarged to 0,80 m at the<br />

supports. The structure is made of concrete<br />

of compressive strength C45/50.<br />

The superstructure is prestressed with<br />

internal bonded tendons (negative in<br />

the cantilever and positive in the span)<br />

3,0%<br />

Drazenci<br />

2<br />

Ptuj<br />

Drava Drava Water<br />

Gravel<br />

Sand/clay<br />

Gravel<br />

Sand/clay<br />

Marl base<br />

18,70<br />

0,50 3,10 1,70<br />

4,05<br />

4,05 1,70 3,10 0,50<br />

3,0%<br />

Drazenci<br />

Ptuj<br />

3,0% 3,0%<br />

Fig. 3: Stay cables anchored adjacent to concrete webs (Units: m)<br />

k.niv.<br />

2,70<br />

and additionally with 2 × 5 pairs of<br />

extrados cables, 31 strands of diameter<br />

15,7 mm. They support the box at<br />

a spacing of 5 m, adjacent to the webs,<br />

so that the force flows directly into the<br />

longitudinal load-carrying system and<br />

represents no problem.<br />

The superstructure also includes short<br />

pylons of total height 8,5 m (L/11,8),<br />

two on each support. The pylons with<br />

the cross section 1,20 m/2,80 m are<br />

inclined outwards with an inclination<br />

7,5 : 1, so that the cables, because of the<br />

route curvature, do not interfere with<br />

the clearance. The deviators for extrados<br />

cables in each pylon are designed<br />

to make possible the replacement of<br />

individual cables. The pylons are vertically<br />

prestressed on the tension side<br />

with dywidag bars with Φ = 40 WR<br />

(950/1050 MPa), with 11 bars on the<br />

inner side of the radius, and with 5 bars<br />

on the outer side. Thus the loading due<br />

to the structure curvature is compensated.<br />

The strength of the concrete for<br />

the pylons is C45/55.<br />

Inclined “Extrados” Cables<br />

In addition to the bonded tendons<br />

(negative and positive) in the girder<br />

and vertical prestressing bars in short<br />

pylons, there is also the third type of<br />

prestressing element, namely, low<br />

inclined stays, also known as extrados<br />

cables. To maintain uniformity, all the<br />

extrados cables are of the same bearing<br />

capacity and with equal number<br />

(31) of bearing strands with the section<br />

150 mm 2 . The cables are 46 to 88<br />

m long with the stressing force from<br />

3600 to 3800 kN.<br />

Each inclined cable consists of the<br />

following:<br />

– the free length;<br />

– two equal anchorages on the girder<br />

beam;<br />

– deviator or saddle located on the<br />

short pylon (Fig. 4).<br />

Free Length<br />

Extrados cables consist of 31 monostrands<br />

of 15,7 mm (cross section<br />

150 mm 2 ) in the outer PE protection<br />

pipe that has been grouted with the<br />

cement mortar.<br />

The composition of the free length of<br />

the inclined cable is as follows:<br />

– Bearing element: seven-wire strand<br />

with nominal diameter of 15,7 mm<br />

and the section 150 mm 2 and of<br />

quality 1570/1770 MPa with a very<br />

low relaxation (


φ193,7<br />

Fig. 4: Short pylons with saddles and vertical prestressing bars<br />

– static test on individual strands;<br />

– dynamic test on individual strands<br />

according to fib recommendations<br />

for extrados cables, which means the<br />

stress modification Δs = 140 MPa<br />

with s k = 0,55f u and in 2 × 10 6 loading<br />

cycles.<br />

The durability of the inclined stays<br />

anchorages is assured by the material<br />

selection and the selected corrosion<br />

protection system. The anchorage head<br />

and the winding nut are made of rustresistant<br />

alloy, all the other anchorage<br />

elements are hot galvanized. The inner<br />

and upper part of the anchorage and<br />

cable are additionally grouted with<br />

cement mortar; the head, wedges and<br />

strands, making possible subsequent<br />

prestressing, are grouted with greases<br />

in the protection cap.<br />

550<br />

550<br />

550/550<br />

After construction, the space between<br />

the monostrands and external<br />

PE cover was grouted with cement<br />

mortar to achieve an additional<br />

mechanical resistance and prevent<br />

water condensation in the pipe.<br />

Anchorage<br />

Each inclined stay has two anchorages<br />

into the girder structure. The typical<br />

anchorage VT 31-150 SK has been<br />

used with a minor modification at the<br />

external head protection (Fig. 5).<br />

6260<br />

Fig. 5: Stay-cable anchorage system (Units: mm)<br />

φ323,6<br />

PEHD φ180<br />

Fig. 6: The saddle consisting of two bent steel tubes (Units: mm)<br />

2200<br />

φ193,7<br />

φ323,6<br />

PEHD φ180<br />

The anchorage and system have been<br />

used and verified on numerous bridges<br />

with inclined stays, having more strict<br />

requirements than those with extrados<br />

cables. In addition, dynamic tests<br />

of the entire anchorage for s k = 0,45f u ,<br />

Δs = 200 MPa and 2 × 10 6 cycles were<br />

carried out.<br />

As the system used had additional<br />

modifications (galvanized wires) especially<br />

for the bridge over the Drava<br />

River, two additional tests were performed,<br />

with positive results:<br />

φ355,6<br />

φ180<br />

Saddle<br />

For anchoring and the transition<br />

through the pylon respectively the<br />

so-called deviating saddle or deviator<br />

was used. This solution is being used at<br />

extrados bridges, as the cable-breaking<br />

angle is smaller than in cable-stayed<br />

bridges. In this case, the curvature<br />

radius of 4,60 m for a length of about<br />

2,2 m was used.<br />

The deviator consists of two bent steel<br />

pipes, namely, of the external pipe<br />

(Φ = 323,6/7,1 mm) and the internal<br />

deviator pipe (Φ = 193,7/5,6 mm).<br />

Within the deviating pipes, the course<br />

of the settled parallel strands was<br />

achieved with PE spacers, assuring distance<br />

between individual strands and<br />

the pipe. Inside the deviation pipe, the<br />

PE cover and the grease were removed<br />

from the monostrands. The prepared<br />

strands were thereafter grouted with a<br />

fast curing acrylic resin mortar, which<br />

is chemically neutral to the galvanized<br />

wires.<br />

The detail of the deviator is shown<br />

in Fig. 6. The external pipes for the<br />

deviators were inserted in a group<br />

into the pylon with the auxiliary steel<br />

structure, assuring a more exact placement<br />

into the inclined pylon, according<br />

to the shop drawings. Owing to the<br />

horizontal curvature of the road and<br />

the vertical curvature of the elevation,<br />

each deviator in the pylon has its own<br />

position.<br />

Construction<br />

The bridge erection, lasting from<br />

October 2005 till May 2007, was a very<br />

Structural Engineering International 4/<strong>2010</strong> Technical Report 465


demanding task. The foundation in<br />

the lake was carried out with artificial<br />

islands that were built with the help of<br />

steel sheet piles. This was followed by<br />

deep foundation with piles of diameter<br />

Φ 1,50 m and of length 25 to 30 m,<br />

reaching the marl base. The transportation<br />

of equipment and material on<br />

the lake was carried out through heavy<br />

barges.<br />

Fig. 7: The balanced cantilever construction with the help of “extrados cables”<br />

Fig. 8: Bridge construction<br />

The bridge deck was built by the balanced<br />

cantilever method with the help<br />

of inclined extrados cables. Movable<br />

scaffolding with segment length of 5 m<br />

was used for erection. Since the superstructure<br />

is of a constant height, the<br />

scaffold was not modified as for the<br />

usual variable height girders. The setting<br />

and prestressing of stay cables was<br />

performed. Doing this saved time, consequently<br />

reducing the duration of a<br />

single construction stage to one week,<br />

which is normal.<br />

One of the particularities of the<br />

construction was the “extrados<br />

cables”, which were assembled near<br />

the bridge and then mounted as a<br />

whole through saddles in the pylon<br />

to anchorages (Fig. 7). Prestressing<br />

was performed with mono jacks from<br />

both sides, using the strand-by-strand<br />

method.<br />

The stressing protocol ensured that<br />

each strand in the stay cable was tensioned<br />

with the same force at the end<br />

of the tensioning procedure. Cables<br />

have an ultimate load capacity of 8200<br />

kN; in service stage being loaded with<br />

approximately 4000 kN.<br />

Fig. 9: Completed bridge<br />

Very complex was also the monitoring<br />

of deformations in construction<br />

stages, due to the flexible cantilevered<br />

deck of the bridge (Figs. 8-11).<br />

The construction of such a bridge<br />

would not have been possible without<br />

the most modern software solutions<br />

in the field of cable-stayed bridges<br />

and the most accurate surveying<br />

equipment.<br />

Conclusion<br />

Fig. 10: Bridge over the Drava River at Ptuj, Slovenia<br />

The solution employing the so<br />

called “extradosed bridge” system<br />

managed to ingeniously overcome<br />

numerous restrictions and difficulties<br />

imposed on the Drava River<br />

spanning. The sharp curvature, which<br />

confronted design and building with<br />

a very challenging task, was successfully<br />

carried out by using some new<br />

design solutions and original structural<br />

details.<br />

466 Technical Report Structural Engineering International 4/<strong>2010</strong>


<strong>SEI</strong> Data Block<br />

Owner:<br />

DARS, d.d., Motorway Company,<br />

Celje, Slovenia<br />

Structural design:<br />

PONTING, d.o.o., Maribor, Slovenia<br />

Contractors:<br />

SCT, d.d., Ljubljana, Slovenia<br />

Porr AG, Wien, Austria<br />

Fig. 11: Night view of the heritage town and its new bridge<br />

The bridge provided the city of Ptuj<br />

a solution for their traffic problems<br />

and a unique engineering structure.<br />

In addition to that, the bridge does<br />

not compete with the city’s urban<br />

cultural heritage; moreover, it has<br />

the potential to become a new architectural<br />

landmark and symbol for the<br />

modern city of Ptuj on its southern<br />

border.<br />

Bridge type: Concrete extrados bridge<br />

Bridge size: L = 433 m,<br />

W = 18,70 m, A = 8097 m 2<br />

Concrete (m 3 ): 9488<br />

Reinforced steel (t): 1300<br />

High quality steel<br />

for cables (kg): 189 900<br />

High quality steel<br />

for stays (kg): 99 060<br />

Total cost (EUR million): 8,8<br />

Service date: May 2007<br />

SED 12 - New Structural Engineering Document Published<br />

IABSE<br />

This SED book provides case studies of structural rehabilitation, repair,<br />

retrofitting, strengthening, and upgrading of structures. Selected studies<br />

are presented in this SED and cover a variety of structural types from<br />

different countries.<br />

This document is a summary of practices to help structural engineers.<br />

The reader will discover different approaches to put forward strengthening<br />

or rehabilitation projects. Even identical technical problems<br />

could have very different efficient solutions, as discussed in the papers,<br />

considering structural, environmental, economic factors, as well as<br />

contractor and designer experience, materials, etc.<br />

Price:<br />

CHF 40 for Members, CHF 70 for Non-Members<br />

www.iabse.org/publications/onlinshop<br />

Structural Engineering Documents<br />

12<br />

Case Studies of<br />

Rehabilitation, Repair,<br />

Retrofitting, and<br />

Strengthening<br />

of Structures<br />

International Association for Bridge and Structural Engineering<br />

Association Internationale des Ponts et Charpentes<br />

Internationale Vereinigung für Brückenbau und Hochbau<br />

IABSE<br />

AIPC<br />

IVBH<br />

Structural SED 12.indd Engineering 1 International 4/<strong>2010</strong> Technical 28/10/10 Report 11:28 467 AM


Recent PhD Abstracts<br />

Structural Engineering International (<strong>SEI</strong>) would like to help you stay up-to-date with some of the exciting and cutting<br />

edge research being carried out at universities around the world. Our new rubric, “Recent PhD Abstracts,” will provide<br />

a window onto recent research activities, while giving recent PhDs a chance to disseminate their results quickly. Readers<br />

wishing to follow up on the information presented in the abstracts will be encouraged to contact the authors, thereby stimulating<br />

direct contact between researchers and those most interested in their results. The full-length manuscript will also<br />

be available using the DOI or the URL that will be given with most of the abstracts. Submission form can be downloaded<br />

from: www.iabse.org/<strong>journal</strong>sei/asanauthor<br />

Simple Models for Analysis of<br />

in-plane Loaded Masonry Walls<br />

Author: Dr. Alvaro Viviescas Jaimes,<br />

Columbia<br />

Email: aviviescasj@hotmail.com<br />

Supervisor: Dr. Pere Roca Fabregat,<br />

Universitat Politècnica de Catalunya,<br />

Barcelona, Spain<br />

URL: www.tesisenxarxa.net/TESIS_<br />

UPC/AVAILABLE/TDX-1229109-<br />

130304//AVJ1de1.pdf<br />

Language of this document: Spanish<br />

A simplified method for the analysis<br />

of the ultimate capacity of walls subjected<br />

to in-plane forces is presented.<br />

The method is based on simple equilibrium<br />

models representing the combination<br />

of compression or tension<br />

stress fields mobilized at the ultimate<br />

condition. The thesis presents tentative<br />

rules for the construction of the<br />

models with some specific models proposed<br />

for walls subjected to different<br />

loading conditions. The performance<br />

of the proposed models is analyzed<br />

by comparison with numerical results<br />

generated by means of the well known<br />

micro-modeling approach specifically<br />

developed for the analysis of masonry<br />

structures.<br />

These simple models are based on the<br />

struts and ties method which uses the<br />

struts to represent the compressive<br />

stress fields and the ties for tension<br />

zones, both of them forming a resistant<br />

mechanism (Fig. 1). The objective<br />

was to predict satisfactorily, by means<br />

of the numerical micro-model, the ultimate<br />

loads and mechanisms of fracture<br />

observed in the experimental tests on<br />

masonry walls. Calibration and validation<br />

by means of experimental results<br />

and numerical simulation constitutes<br />

an essential and necessary feature of<br />

the method.<br />

The examples of application presented,<br />

corresponding to walls subjected to<br />

uniform vertical loading or confined<br />

walls subjected to uneven vertical loading,<br />

illustrate the ability of the models<br />

Numerical model<br />

Fig. 1: Numerical and Simple Model<br />

to satisfactorily predict the maximum<br />

horizontal forces that the walls can<br />

resist. More specifically, the models<br />

have been able to provide good estimations<br />

of reference strength values<br />

obtained by means of an up-to-date<br />

micro- model. The parametric studies<br />

carried out show that the simple models<br />

proposed take into account adequately<br />

the influence of the geometry<br />

(in particular, the width to height ratio)<br />

and the main material properties (the<br />

masonry compression strength and<br />

joint-unit frictional properties) on the<br />

in-plane strength of the wall.<br />

Railway Bridge Response to<br />

Passing Trains: Measurements<br />

and FE Model Updating<br />

Author: Dr. Johan Wiberg, Sweden<br />

Email: jw@kth.se<br />

Supervisor: Prof. Raid Karoumi,<br />

Prof. Håkan Sundquist; KTH Royal<br />

Institute of Technology, Dept. of Civil<br />

and Architectural Eng., Division of<br />

Structural Design and Bridges,<br />

Stockholm, Sweden<br />

URL: http://kth.diva-portal.org/smash/<br />

get/diva2:241373/FULLTEXT02<br />

Language of this document: English<br />

Existing railway bridges are being<br />

analysed in detail for their response<br />

a<br />

r c<br />

b c<br />

x f<br />

Simple model<br />

to moving loads due to the increase in<br />

speeds and axle loads. These numerical<br />

analyses are time consuming as<br />

they involve many simulations using<br />

different train configurations at different<br />

speeds as well as many other<br />

considerations. Thus, simplified models<br />

are often chosen for practical and<br />

time efficient simulations. The New<br />

Årsta Railway Bridge in Stockholm<br />

was successfully instrumented during<br />

construction and a simplified 3D beam<br />

element FE model was prepared. The<br />

model was first manually tuned based<br />

on static load testing. The most extensive<br />

work was performed in a statistical<br />

identification of significantly<br />

influencing modelling parameters to<br />

be included in an optimised FE model<br />

updating. The amount of parameters<br />

included in the optimisation was in<br />

this way kept at an optimally low level.<br />

For verification, measurements from<br />

several static and dynamic field tests<br />

with a fully loaded macadam train and<br />

Swedish Rc6 locomotives were used.<br />

The implemented algorithms were<br />

shown to operate efficiently and the<br />

accuracy in static and dynamic load<br />

effect predictions was considerably<br />

improved. It was concluded that the<br />

complex bridge can be simplified by<br />

means of beam theory and an equivalent<br />

modulus of elasticity for simplified<br />

global analyses. That modulus was<br />

b p<br />

468 Recent PhD Abstracts Structural Engineering International 4/<strong>2010</strong>


in this case approximately 25% larger<br />

than the specified mean value for the<br />

concrete grade in question. The optimised<br />

FE model was used in moving<br />

load simulations with high speed train<br />

loads according to the design codes.<br />

Typically, the calculated vertical acceleration<br />

of the bridge deck was lower<br />

than the allowable code value. This<br />

indicates that multispan continuous<br />

concrete bridges are not so sensitive<br />

to train induced vibrations and may be<br />

suitable for high speed traffic. Finally,<br />

the relevant area of introducing the<br />

proposed FE model updating procedure<br />

in the early bridge design phase<br />

is outlined.<br />

Residual Sresses in Stainless<br />

Steel Box Sections<br />

Author: Dr. Michal Jandera, Czech<br />

Republic<br />

Email: michal.jandera@fsv.cvut.cz<br />

Supervisor: Prof. Ing. Josef Machacek,<br />

Dr. Sc, Czech Technical University,<br />

Prague<br />

URL: www.ocel-drevo.fsv.cvut.cz/<br />

ODK/cz/docs/Disertace/Disertace-<br />

Jandera.pdf<br />

Language of this document: Czech<br />

The investigation is focused on stainless<br />

steel cold rolled SHS. In total, 14<br />

SHS stainless steel stub columns were<br />

tested for subsequent FEM validation.<br />

The material properties of the flat and<br />

corner areas of the sections as well as<br />

the initial deflections of all plates were<br />

investigated and compared with existing<br />

predictive formulae. The research<br />

embraces experimental investigation<br />

of residual stresses induced by the<br />

forming process of the sections. The<br />

residual stress pattern in the sections<br />

was determined according to the sectioning<br />

method for the longitudinal<br />

and as well as the transversal direction.<br />

In addition, trough-thickness<br />

measurements using an X-ray diffraction<br />

method were performed. Patterns<br />

of the membrane and bending stress<br />

distribution along the section were<br />

generalized and suitable predictive<br />

formulas for general use were developed.<br />

A high correlation between the<br />

suggested formulas and measured<br />

pattern is demonstrated. The patterns<br />

were finally used in FEM parametric<br />

studies to show the possible influences<br />

of each part of the residual stresses on<br />

the compressive strength of sections<br />

subjected to local and global buckling.<br />

The studies are assessed for all substantial<br />

parameters such as web plate<br />

slenderness, column slenderness and<br />

also for the nonlinearity parameter of<br />

the Ramberg-Osgood formula. Paradoxically,<br />

it was found that inclusion<br />

of residual stresses in stainless steels<br />

(unlike common carbon steels) generally<br />

led to an increase in the loadcarrying<br />

capacity. This was attributed<br />

principally to the influence of the<br />

bending residual stresses on the material<br />

stress–strain curve. It was found<br />

that despite the secant modulus being<br />

consistently reduced in the presence<br />

of the residual stresses, the tangent<br />

modulus was increased in some<br />

regions of the stress–strain curve. For<br />

cases where column failure strains<br />

coincided with these increased tangent<br />

modulus regions (which was over the<br />

majority of the practical slenderness<br />

range, except large column slenderness),<br />

higher buckling loads resulted.<br />

Design Oriented Constitutive<br />

Model for Steel Fiber<br />

Reinforced Concrete<br />

Author: Dr. Filipe Laranjeira de<br />

Oliveira, Spain<br />

Email: filipe.laranjeira@gmail.com<br />

Supervisor: Prof. Antonio Aguado,<br />

Prof. Climent Molins, Universitat<br />

Politècnica de Catalunya<br />

URL: www.tesisenxarxa.net/TDX-<br />

0602110-115910/<br />

Language of this document: English<br />

In the last years, industry has been<br />

demanding the use of steel fiber reinforced<br />

concrete (SFRC) in structural<br />

applications (Fig. 2). Because the postcracking<br />

strength of this material is not<br />

negligible, the crack-bridging capacity<br />

provided by fibers may replace,<br />

partially or completely, conventional<br />

steel reinforcement. Therefore, an<br />

appropriate characterization of the<br />

SFRC uniaxial tensile behavior is of<br />

paramount interest. However, in spite<br />

Fig. 2: Uniaxial tensile test on concrete<br />

specimen with 60kg/m 3 of steel fiber<br />

reinforcement<br />

of the extensive research and recently<br />

advanced standards, there is no agreement<br />

on the constitutive model to<br />

be used for the design of SFRC. The<br />

crack-bridging capacity provided by<br />

steel fibers improves both the toughness<br />

and the durability of concrete.<br />

Conventional SFRC is a material that<br />

presents a softening response under<br />

uniaxial tension, but may develop<br />

hardening behavior in bending due to<br />

its ability to redistribute stresses within<br />

the cross-section. This evidence has<br />

contributed to an increasing interest<br />

and growing number of applications<br />

of this material. In this doctoral thesis,<br />

a direct and rationale approach to predict<br />

the tensile response of SFRC for<br />

structural design calculations is developed.<br />

The proposed design-oriented<br />

constitutive model differentiates itself<br />

from previous studies in multiple<br />

aspects and defines a new philosophy<br />

for the design of SFRC elements. This<br />

model provides a direct and practical<br />

procedure to obtain the material’s<br />

tensile behavior by means of parameters<br />

with physical meaning and based<br />

on clear concepts: fiber pullouts and<br />

orientations. One of the major contributions<br />

of this work is the ability to<br />

predict the stress-crack width curves<br />

that reflect the specific combination of<br />

the properties of the matrix and fibers<br />

applied. Furthermore, it introduces<br />

a novel philosophy for the material<br />

design by taking into account influences<br />

from the production process,<br />

fresh-state properties and the element<br />

to be built in order to define the constitutive<br />

diagram.<br />

Structural Engineering International 4/<strong>2010</strong> Recent PhD Abstracts 469


Eminent Structural Engineer: Christian Menn—Bridge<br />

Designer and Builder<br />

Eugen Brühwiler, Prof., Dr. Civil Eng., ETH, Ecole Polytechnique Fédérale de Lausanne (EPFL), Lausanne, Switzerland.<br />

Contact: eugen.bruehwiler@epfl.ch<br />

Brief CV<br />

1927 Born on 3 March in<br />

Meiringen, Switzerland.<br />

1950 Graduated from the<br />

Swiss Federal Institute<br />

of Technology (ETH)<br />

Zurich with a degree in<br />

Civil Engineering.<br />

1953 Accepted a position as<br />

assistant to Professor<br />

Pierre Lardy at ETH<br />

Zurich, where he<br />

completed his doctoral<br />

degree in Civil<br />

Engineering in 1956.<br />

1956 Worked as an engineer<br />

with the French<br />

construction contractor<br />

Dumez, on the<br />

construction of the<br />

UNESCO building in<br />

Paris.<br />

1957 Established his own<br />

engineering design<br />

office, specializing in<br />

reinforced concrete<br />

construction, in Chur,<br />

Switzerland.<br />

1959 Built his first bridges at<br />

Letziwald and Cröt in<br />

Switzerland.<br />

1971–1992 Was Professor of<br />

Structural Engineering<br />

at the Swiss Federal<br />

Institute of Technology<br />

(ETH) Zurich.<br />

Since 1992 Commenced private<br />

practice as a consulting<br />

engineer.<br />

1996 Received honorary<br />

doctorate from the<br />

University of Stuttgart.<br />

2008 Received honorary<br />

doctorate from the Ecole<br />

Polytechnique Fédérale<br />

de Lausanne (EPFL).<br />

2009 Received the<br />

International Award<br />

of Merit in Structural<br />

Engineering from IABSE.<br />

Fig. 1: Christian Menn<br />

Introduction<br />

Christian Menn (Fig. 1) ranks among<br />

the most important and creative bridge<br />

design engineers of the recent decades.<br />

His bridges are witnesses of his exclusive<br />

engagement in bridge engineering<br />

and more than 50 years of continuous<br />

experimentation in conceptual,<br />

structural and aesthetic design. They<br />

teach us that the true art of structural<br />

engineering is characterized by<br />

innovation and imagination with the<br />

objective to improve the environment<br />

through structural art. His work has<br />

been exhibited at various universities<br />

in Europe and at art museums<br />

across the United States, and attests to<br />

his extraordinary creativity and absolute<br />

mastery of bridge engineering.<br />

Fig. 3: Reichenau bridge at Tamins, Switzerland, 1963<br />

Bridges in Switzerland<br />

The first bridges designed by him<br />

(Fig. 2) were inspired by the innovative<br />

bridge forms—the three-hinged hollow<br />

box girder and the deck-stiffened<br />

arch—of his Swiss predecessor Robert<br />

Maillart. After 1960, Menn developed<br />

his own style of the arch bridge. This<br />

consists of a monolithic frame system<br />

composed of a thin polygonal<br />

arch supporting slender cross walls<br />

at a wide spacing, and a stiff partially<br />

prestressed box girder deck with a<br />

wide cantilevered roadway. The most<br />

prominent example is the 100 m span<br />

Reichenau Bridge (Fig. 3) over the<br />

Rhine River, which represents a new<br />

form for deck-stiffened arch bridges<br />

with long spans. Menn has used this<br />

Fig. 2: Letziwald bridge at Avers, Switzerland,<br />

1959<br />

470 Eminent Structural Engineer Structural Engineering International 4/<strong>2010</strong>


system for several other arch bridges,<br />

such as the Viamala Bridge and the<br />

Nanin and Cascella bridges on the<br />

Moesa River, built between 1966 and<br />

1968 in the canton of Graubünden.<br />

In 1970, in collaboration with an engineering<br />

firm, Menn won a competition<br />

to build the six-lane 1100 m long<br />

Felsenau highway bridge across the<br />

Aar Valley in Bern. He designed a<br />

structure in prestressed concrete with<br />

two central spans of 156 m each. The<br />

hollow box girder with inclined webs<br />

and variable depth is curved and supports<br />

a 7,6 m wide cantilevered deck<br />

slab. The bridge is characterized by<br />

its slender appearance and impressive<br />

aesthetic despite its large size (Fig. 4).<br />

In 1971, Menn accepted a position as<br />

professor at ETH Zurich. He left his<br />

engineering office but continued to<br />

design exceptional bridges throughout<br />

Switzerland in collaboration with<br />

different engineering firms. He regularly<br />

sat on juries for bridge design<br />

competitions.<br />

Christian Menn designed the Ganter<br />

Bridge built in 1980 on the Simplon<br />

Pass road, whose extraordinary silhouette<br />

presents a new aesthetic for<br />

Fig. 4: Felsenau bridge at Berne, Switzerland,<br />

1975<br />

Fig. 5: Ganter bridge at Eisten, Switzerland, 1980<br />

structures in the Swiss Alpine environment<br />

(Fig. 5). The structure consists of<br />

steel cable stays encased in a sheet of<br />

concrete and anchored in short pylons<br />

to form a rigid structure 150 m above<br />

the valley floor. The Ganter’s central<br />

span of 174 m is still the longest in<br />

Switzerland. The distinctive structural<br />

form arose from Menn’s imagination<br />

and aesthetic inspirations.<br />

In the same period, Menn collaborated<br />

with an engineering firm in Bellinzona<br />

to win a design competition for the<br />

Biaschina Viaduct, a highway bridge<br />

completed in 1983 in Switzerland’s<br />

Ticino valley. The bridge’s tall columns<br />

and roadway, built using a balanced<br />

cantilever construction method, result<br />

in a form of harmonious proportions.<br />

Towards the end of the 1980s, in collaboration<br />

with another engineering<br />

firm, Menn conceived the Chandoline<br />

Bridge, a cable-stayed structure over<br />

the Rhone River at Sion.<br />

At the end of his academic career,<br />

Menn designed the Sunniberg Bridge,<br />

built between 1996 and 1998 to divert<br />

highway traffic around the town of<br />

Klosters in the canton of Graubünden<br />

(Fig. 6). This five-span cable-stayed<br />

structure is 526 m long and crosses the<br />

valley at a height of 50 to 60 m. The<br />

roadway’s strong curvature in the plan<br />

required that the pylons lean away<br />

from the bridge deck so that the cables<br />

did not interfere with the road clearance.<br />

Because of its strong curvature,<br />

the bridge deck could be fixed to the<br />

abutments without dilation joints, and<br />

longitudinal deck length variation is<br />

taken by radial displacements of the<br />

structural system. This is probably the<br />

first time this solution is applied worldwide.<br />

The slender pylons, thin deck, and<br />

harped cables form an elegant ensemble<br />

and confer a striking aesthetic<br />

on the bridge, expressing boldness in<br />

and inviting admiration for the art of<br />

engineering. In 2001, the Sunniberg<br />

Fig. 6: Sunniberg bridge at Klosters, 1998<br />

Bridge was awarded the Outstanding<br />

Structure Award by the International<br />

Association for Bridge and Structural<br />

Engineering (IABSE).<br />

Bridges Worldwide<br />

After retiring from professorship in<br />

1992, Menn began a new vocation as<br />

consultant for the design of several<br />

significant bridges around the world.<br />

This pursuit has allowed him to further<br />

develop his passion for the art of<br />

bridge design. A series of innovative<br />

bridge projects arose from situations<br />

calling for structures of both functional<br />

and symbolic importance, notably<br />

in the United States.<br />

In 1991, Menn proposed a cable-stayed<br />

bridge to the civic leaders of Boston as<br />

a means to satisfy several complex project<br />

constraints. The bridge’s original<br />

character stems from the play in the spatial<br />

configuration of stay cables as experienced<br />

by automobilists crossing the<br />

bridge (Fig. 7). The cables on the 227 m<br />

central span are anchored to the outer<br />

edges of the bridge deck—at 60 m wide,<br />

it is the widest cable-stayed bridge in<br />

the world—while on the end spans, the<br />

cables are anchored along the median.<br />

The two “inverted Y-shaped” towers<br />

symbolize the entrance into downtown<br />

Boston. The stay cable arrangement<br />

allows for a significant reduction in<br />

transverse bending of the pylons due<br />

to the asymmetric cross section and<br />

eccentric traffic loads. The concept of<br />

the bridge structure is the product of<br />

a design process led by solely optimizing<br />

the flow of forces while respecting<br />

the stringent boundary conditions.<br />

The aesthetic appearance results thus<br />

simply from an optimized and refined<br />

structural form. The Leonard P. Zakim<br />

Bunker Hill Bridge, named after a civil<br />

rights activist, opened on 12 May 2002<br />

with more than 200 000 people crossing<br />

the bridge on foot. The bridge has<br />

received several distinctions, and in<br />

honour of the designer, 3 November<br />

2000 was proclaimed as “Christian<br />

Menn Day” in Massachusetts.<br />

Structural Engineering International 4/<strong>2010</strong> Eminent Structural Engineer 471


Future Bridge Project<br />

Fig. 7: Leonard P. Zakim Bunker Hill Bridge in Boston, USA, 2002<br />

Menn developed the designs of several<br />

other major bridges in the United<br />

States. He designed the winning proposal<br />

for a bridge competition held in<br />

1998 for the 2 km long Woodrow Wilson<br />

Bridge crossing the Potomac River,<br />

south of Washington, DC, but only the<br />

basic structural system remained from<br />

Menn’s original design. The originally<br />

slender, elegant V-shaped piers were<br />

abandoned in favor of unmotivated<br />

curved massive elements.<br />

Since 2001, Menn has developed several<br />

designs for a new bridge crossing<br />

the Niagara River near its famous falls.<br />

These designs for the “Peace Bridge”<br />

between Buffalo and Fort Erie have<br />

won the support largely of the public<br />

and local leaders, but the official process<br />

of approvals is still ongoing.<br />

Fig. 8: Streicker Pedestrian Bridge over<br />

Washington Road on the Princeton<br />

University Campus, Princeton, USA, 2009<br />

Fig. 9: Grimsel lake bridge at Guttannen, Switzerland (Project)<br />

Menn’s conceptual designs for a bridge<br />

crossing the valley in front of Hoover<br />

Dam near Las Vegas, the east span of<br />

the San Francisco Oakland Bay Bridge,<br />

a double self-anchored suspension<br />

bridge for crossing the Ohio River at<br />

Louisville and a new Inner Belt Bridge<br />

in Cleveland were unfortunately all<br />

not considered, sometimes because of<br />

their seemingly too innovative nature,<br />

and more conservative approaches<br />

were preferred instead.<br />

On the Princeton University campus,<br />

Menn’s design of the slender pedestrian<br />

bridge passing over Washington<br />

Road took the form of a bold arch. The<br />

Streicker Bridge, completed in <strong>2010</strong> and<br />

named after its principal benefactor,<br />

links four buildings designed by internationally<br />

acclaimed architects (Fig. 8).<br />

In 2007, Christian Menn was asked<br />

to propose bridge designs to provide<br />

access to a large island to be developed<br />

in Abu Dhabi in the United Arab<br />

Emirates. For the two largest most<br />

em blematic bridges, he proposed an<br />

arch rising above the roadway surface<br />

from which the bridge deck is suspended,<br />

and a cable-stayed bridge with a single<br />

“spindle-shaped” pylon supporting a<br />

wide, softly curving roadway just a few<br />

meters above the water surface. These<br />

projects are currently in the design<br />

development phase, and construction is<br />

expected to begin in the near future.<br />

In Switzerland, near the birthplace<br />

of Menn, plans to raise the Grimsel<br />

dam and lake require redrawing of<br />

the road over the Grimsel Pass which<br />

has to cross more than 300 m across<br />

the artificial lake. In 2005, Christian<br />

Menn designed a cable-stayed bridge<br />

crossing the lake with a single span<br />

of 352 m. This will be the longest in<br />

Switzerland, supported by two vertical<br />

75-m high inversed Y-shaped pylons.<br />

The aesthetic expression is characterized<br />

by the slender deck and purely<br />

shaped pylons, as well as the semifan<br />

arrangement of the stay cables<br />

and their concrete anchorage blocks<br />

shaped like mountain crystals (Fig. 9).<br />

Lessons in Bridge Design<br />

Christian Menn has designed arch, box<br />

girder and cable-stayed bridge structures.<br />

He has developed new ideas for<br />

each of these types of structural systems<br />

in order to create unique forms<br />

and an original aesthetic. His bridges<br />

are the result of a design process reducing<br />

structural elements to meet the<br />

given functional and environmental<br />

requirements. They are characterized<br />

by harmonious proportioning of structural<br />

elements, a slender appearance<br />

and coherent integration of shapes,<br />

which continually led to new creations<br />

and expressions. Menn’s works are<br />

among the most beautiful bridges built<br />

in the last 50 years; they are symbols of<br />

modern technology, and they establish<br />

a technical aesthetic.<br />

Menn’s bridges express technical efficiency<br />

with an accent on slenderness<br />

and transparency. They emphasize to us<br />

the importance of understanding how<br />

structural systems function. A sound<br />

engineering concept is the solid basis<br />

for a far-reaching aesthetic quality and<br />

for finding simple yet elegant structures.<br />

Guided by the basics of structural<br />

mechanics and natural sciences, this<br />

approach continues to be very efficient<br />

and valuable, in particular nowadays,<br />

when (architect-led) bridge designs,<br />

often based on a spectacular metaphoric<br />

idea rather than on an efficient structural<br />

concept, have produced structures<br />

that are excessively expensive to build<br />

and maintain and thus are controversial.<br />

In his numerous lectures on bridge<br />

design, Christian Menn has always<br />

insisted that the art of structural engineering<br />

should be appreciated and be<br />

given much more importance, in particular,<br />

in bridge design competitions and in<br />

the education of structural engineers.<br />

472 Eminent Structural Engineer Structural Engineering International 4/<strong>2010</strong>


IABSE Annual Meetings<br />

Venice, Italy, September 19–21, <strong>2010</strong><br />

<strong>SEI</strong> Editorial Board and Nominees<br />

The <strong>2010</strong> Annual Meetings were held<br />

at the Palazzo del Casino, at the Lido<br />

in Venice, prior to the 34 th IABSE<br />

Symposium. There was a good attendance<br />

with 140 participants and 55<br />

accompanying persons. Accompanying<br />

persons discovered Venice and islands<br />

and enjoyed the social events together<br />

with committee members: on Sunday<br />

Ms Evelyne Stampfli, Deputy Consul<br />

General of Switzerland, gave a reception<br />

at the Excelsior Hotel. On Monday<br />

evening, Jacques Combault, President<br />

of IABSE welcomed all delegates, and<br />

Carlo Urbano, Chair of the Italian<br />

Group, in his turn welcomed all to<br />

Venice with drinks and delicious food<br />

on Tuesday evening.<br />

The Annual Meetings included the<br />

Administrative, Executive, Permanent<br />

Committee and Chair National Groups,<br />

Technical Committee, Editorial, Correspondents<br />

and E-Learning Boards,<br />

Working Commissions, Work ing<br />

Groups, Scientific Committees, Young<br />

Engineers Board, the Advisory Group<br />

to the Executive Committee and the<br />

IABSE Foundation Council.<br />

The Permanent Committee approved<br />

the annual statement of accounts<br />

2009 and the budget 2011. The annual<br />

accounts 2009 closed with total net<br />

revenues of CHF 989'066 and an<br />

excess of revenues of CHF 17'505. The<br />

Association funds amount to CHF<br />

467'658 as on December 31, 2009.<br />

For the year 2011 a budget with total<br />

net revenues of CHF 1'087'700 was<br />

approved.<br />

The Permanent Committee changed<br />

the article 14 of the By-Laws and made<br />

English the sole official language of<br />

IABSE as from January 1, 2011.<br />

Predrag (Pete) Popovic, USA, New President of IABSE<br />

Jacques Combault ended his term as<br />

President on October 31, <strong>2010</strong>. At the<br />

Closing Ceremony of the 34 th IABSE<br />

Symposium he thanked his colleagues<br />

on the Administrative and Executive<br />

Committees, and all those who have<br />

made the Technical Committee more<br />

dynamic and efficient by encouraging<br />

the creation of new Working Groups<br />

and making IABSE E-Learning<br />

become a reality. During the three<br />

years of his presidency several successful<br />

international events were held:<br />

the Congress in Chicago, Symposia in<br />

Bangkok and Conferences in Helsinki<br />

and Dubrovnik, a Bridge Workshop<br />

and spectacular tour in China. Jacques<br />

Combault attended four Outstanding<br />

Structure Award plaque presentations<br />

and welcomed one new National<br />

Group to the Association. The 80<br />

years of IABSE were celebrated by<br />

making all IABSE publications from<br />

1929–1999 available for free online<br />

to the general public. A new more<br />

dynamic website is on its way to serve<br />

the Association for an even better<br />

exchange of structural engineering<br />

knowledge.<br />

Pete Popovic took the opportunity<br />

during the Closing Ceremony to thank<br />

Jacques Combault and his wife Danièle,<br />

who has assisted and supported her<br />

husband during his presidency.<br />

Pete Popovic from Wiss, Janney,<br />

Elstener Associates, Inc., USA, has<br />

taken office as President of IABSE<br />

on November 1, <strong>2010</strong>, for a period of<br />

three years. He is the second IABSE<br />

President from USA. Pete Popovic<br />

is Member of IABSE since 1985 and<br />

knows the Association well. He has<br />

been Chair of Working Commission 8,<br />

Member of the Technical Committee,<br />

Member of the Outstanding Structure<br />

Pete Popovic, USA, President of IABSE<br />

Award Committee and Vice-President<br />

of IABSE. His contributions to IABSE<br />

Structural Engineering International 4/<strong>2010</strong> Panorama 473


conferences are numerous and he was<br />

the Chair of the Organising Committee<br />

for the Chicago Congress in 2008.<br />

Pete Popovic’s fields of expertise are<br />

the design, assessment and repairs of<br />

bridges and buildings. He has in particular<br />

expertise in assessment and repair<br />

of concrete structures and of fatigue<br />

damage in steel bridges, and exterior<br />

facades of high-rise buildings.<br />

During the first ten years of practice,<br />

he participated in structural design<br />

of major steel bridges and rapid transit<br />

systems in Chicago, New York and<br />

Atlanta, USA. He was engaged in the<br />

design of post-tensioned box girder<br />

bridges in Kuwait. Over the last 30<br />

years, he has evaluated and designed<br />

repairs for over 1500 structures. Major<br />

projects included assessment of steel<br />

bridges for fatigue damage, investigation<br />

of collapses of bridges and buildings,<br />

assessment and design of repairs<br />

for exterior facades of high-rise buildings<br />

up to 60-stories tall, and assessment<br />

and repair of over 100 parking<br />

structures.<br />

Pete Popovic has published over 40<br />

technical papers on assessment, load<br />

testing, strengthening and repair of<br />

bridges, buildings and parking structures<br />

and is a contributing author to<br />

Jacques and Danièle Combault and Pete Popovic<br />

several books. He has received awards<br />

from the International Concrete<br />

Repair Institute for innovative repair<br />

projects and is an invited lecturer at<br />

the University of Wisconsin, USA and<br />

World of Concrete (USA and Mexico)<br />

on topics of concrete repairs, rehabilitation<br />

of parking structures, and prevention<br />

of structural failures.<br />

As President of IABSE Pete Popovic<br />

intends, to work in making IABSE<br />

more visible and increasing IABSE<br />

membership. His goal is to have<br />

National Groups play an increasing<br />

role in recruiting new members and<br />

retaining existing members. The goal is<br />

to increase IABSE membership by 500<br />

over the next three years.<br />

IABSE Awards <strong>2010</strong><br />

Jacques Combault, President of IABSE,<br />

presented the IABSE Awards at the<br />

Permanent Committee (Honorary<br />

Memberships) and at the Opening<br />

Ceremony of the 34 th IABSE Symposium<br />

in Venice on the 21 st and 22 nd of<br />

September.<br />

Honorary Membership<br />

Honorary Membership is presented to<br />

an Individual Member of IABSE, for<br />

exceptionally great services rendered<br />

to the Association.<br />

The Executive Committee of IABSE<br />

has awarded Honorary Membership<br />

to Prof. Aarne Jutila, Finland. The<br />

President of IABSE presented the<br />

Award at the Permanent Committee<br />

meeting on September 21, <strong>2010</strong>,<br />

‘in recognition to his outstanding<br />

and dedicated services to the<br />

Association’.<br />

Aarne Jutila, Finland<br />

Born 1940 in Helsinki, Aarne Jutila<br />

received his Civil Engineering degree<br />

at Helsinki University of Technology<br />

(TKK) in 1966, with major subject<br />

“Bridge Engineering”. After<br />

graduation he studied a year at ETH,<br />

Zurich, as “Bundesstipendiat” under<br />

the guidance of Bruno Thürlimann.<br />

Later he worked as bridge designer<br />

at Kjessler and Mannerstråle AB<br />

in Stockholm and Tapiola, Finland,<br />

as Assistant Lecturer at Queen’s<br />

University of Belfast, Northern Ireland,<br />

and as section chief at the Finnish<br />

Road Administration’s bridge design<br />

office (TVH) in Helsinki before founding<br />

of and working for three consulting<br />

engineering companies. Besides that<br />

he also worked as assistant, laboratory<br />

engineer and, since 1984, as Professor<br />

of Bridge Engineering at TKK. He<br />

retired in August 2008 and continues<br />

his bridge engineering activity as<br />

Managing Director of Extraplan Oy,<br />

consulting engineers, that he founded<br />

in 1977.<br />

Aarne Jutila joined IABSE in 1967,<br />

and has since then held numerous<br />

functions within the Association:<br />

Secretary of the Finnish Group<br />

1972–88 and Chair since that, Vice-<br />

Chair of the Organising Committee<br />

of the 1988 Helsinki Congress, <strong>SEI</strong><br />

474 Panorama Structural Engineering International 4/<strong>2010</strong>


Correspondent and Member of the<br />

Editorial Board 1991–2000, Member of<br />

several scientific committees (Malmö<br />

1999, New Delhi 2005, Dubrovnik<br />

<strong>2010</strong>), Chair of the Scientific<br />

Committee of the Lahti Conference<br />

in 2001, Member of the Executive<br />

Committee and Vice-President<br />

1999–2007.<br />

He continues his engagement for<br />

IABSE: Chair of the Finnish Group,<br />

Member of the Permanent Committee,<br />

Member of the Foundation Council<br />

Board, Member of the Advisory<br />

Group to the Executive Committee of<br />

IABSE.<br />

Honorary Membership<br />

Hai-Fan Xiang, China<br />

The Executive Committee of IABSE<br />

has awarded Honorary Membership<br />

to Prof. Hai-Fan Xiang, China. Jacques<br />

Combault gave a speech at the<br />

Permanent Committee meeting on<br />

September 21, <strong>2010</strong>, and informed that<br />

Prof. Xiang, was not able to travel to<br />

Venice and that Pete Popovic, future<br />

President of IABSE, would present the<br />

Diploma to Prof. Xiang at a Ceremony<br />

at Tongji University, ‘in recognition of<br />

his outstanding and dedicated services<br />

to the Association’.<br />

Hai-Fan Xiang graduated from Tongji<br />

University in 1955 and acquired<br />

his master in 1958. Since then he<br />

has worked at Tongji University for<br />

more than 50 years. He gained the<br />

Research Fellowship of Alexander von<br />

Humboldt Foundation and worked as<br />

a visiting professor at Ruhr University,<br />

Bochum, Germany in 1981 and 1982.<br />

As a pioneer in bridge wind engineering<br />

in China, he devoted his research<br />

field to the wind-resistance of longspan<br />

bridges after returning to Tongji<br />

University in 1982. He was awarded<br />

the title of National Outstanding<br />

Expert in 1986. In 1995, he was elected<br />

as an Academician of the Chinese<br />

Academy of Engineering. In 2007,<br />

he became an Emeritus Professor of<br />

Tongji University.<br />

He has been the first Chairman of<br />

Department of Bridge Engineering,<br />

the founding Dean of College of Civil<br />

Engineering, and the Director of the<br />

State Key Laboratory for Disaster<br />

Reduction in Civil Engineering.<br />

He has published 12 books, numerous<br />

articles domestically and internationally.<br />

He has received more<br />

than 20 national awards and several<br />

international awards including<br />

the R.H. Robert Scanlan Medal<br />

of ASCE and the Anton Tedesko<br />

Medal of the IABSE Foundation<br />

for the Advancement of Structural<br />

Engineering.<br />

Hai-Fan Xiang is President of the<br />

Insti tution of Bridge and Structural<br />

Engineering of China, and Chairman<br />

or Co-Chairman for more than ten<br />

organisations. He joined IABSE in<br />

1992 and has dedicated his time to<br />

the Association on several Scientific<br />

Committees for conferences held in<br />

Seoul 2004, Shanghai 2004 (Chair)<br />

and New Delhi 2005. Former Vice-<br />

President of IABSE (2001-2009),<br />

he is currently a Delegate to the<br />

Permanent Committee and on the<br />

Structural Engineering International<br />

(<strong>SEI</strong>) Advisory Board and a Member<br />

of the Foundation Council of<br />

IABSE.<br />

Dagu Bridge, Tianjin, China<br />

International Award of Merit in<br />

Structural Engineering<br />

The International Award of Merit in<br />

Structural Engineering is conferred<br />

for outstanding contributions in the<br />

field of structural engineering, with<br />

special reference to their usefulness<br />

to society. Contributions may include<br />

various aspects in Planning, Design,<br />

Construction, Materials, Equipment,<br />

Education, Research, Government,<br />

and Management. The Executive<br />

Committee of IABSE has conferred<br />

the International Award of Merit in<br />

Structural Engineering to Man-Chun<br />

Tang, USA, “for blending art and<br />

engineering, and together successfully<br />

creating innovative concepts for signature<br />

bridges that are admired by<br />

both his peers and the general public<br />

alike”.<br />

Man-Chung Tang, USA<br />

Man-Chung Tang is the Technical<br />

Director and Chairman of the Board<br />

Structural Engineering International 4/<strong>2010</strong> Panorama 475


of T.Y. Lin International, a globally<br />

recognised consulting firm with headquarters<br />

in San Francisco, USA.<br />

Man-Chung Tang received his Doctor<br />

of Civil Engineering in 1965 from<br />

the Technical University Darmstadt,<br />

Germany. His career spans more than<br />

44 years, and encompasses designing<br />

and constructing over 100 bridges<br />

worldwide, including 32 cable-stayed<br />

bridges, four major suspension bridges,<br />

and numerous segmental bridges. A<br />

true leader and icon, Dr. Tang’s contributions<br />

to innovations in bridge<br />

design are demonstrated through<br />

teaching, writing over 100 technical<br />

papers, and offering numerous presentations.<br />

He is an honorary professor at<br />

ten universities, a member of the U.S.<br />

National Academy of Engineering, a<br />

foreign member of Chinese Academy<br />

of Engineering, and an honorary member<br />

of the American Society of Civil<br />

Engineers (ASCE).<br />

A world authority on cable-stayed<br />

bridges, Man-Chung Tang served as<br />

Chairman of the American Society of<br />

Civil Engineers (ASCE) committee on<br />

cable-suspended bridges and published<br />

the definitive guideline for the design<br />

of cable-stayed bridges, used today by<br />

engineers all over the world. Dr. Tang<br />

is also a founding member of the Post-<br />

Tensioning Institute (PTI) committee<br />

that published “Recommendations for<br />

the Design and Testing of Stay Cables,”<br />

also used worldwide.<br />

Man-Chung Tang’s bridges, besides<br />

being safe, functional and economical,<br />

are considered works of art–beautiful<br />

structures that blend seamlessly with<br />

their surroundings. It is often quoted<br />

that “the sun never sets on a Dr. Tang<br />

bridge,” as his designs can be found<br />

all around the globe. Man-Chung<br />

Tang continues to advance the field<br />

of bridge engineering as an innovator<br />

and an educator and he has been continually<br />

recognised by his peers for his<br />

dedication to the field.<br />

IABSE Prize<br />

The IABSE Prize was established in<br />

1982 to honour a Member early in<br />

his, or her career for an outstanding<br />

achievement in the field of structural<br />

engineering, in Research, Design or<br />

Construction. The Prize is presented to<br />

Individual Members of IABSE, forty<br />

years of age or younger.<br />

The Executive Committee of IABSE<br />

has presented the IABSE Award <strong>2010</strong><br />

to Roberto Revilla Angulo, Spain, ‘in<br />

Roberto Revilla Angulo, Spain<br />

recognition of his significant involvement<br />

in many major bridge projects,<br />

specially for his contribution in the<br />

design, project of Montabliz Viaduct’.<br />

Roberto Revilla Angulo, was born in<br />

Bilbao (Spain) in 1970. He studied at the<br />

Technical Civil Engineers University<br />

College of Santander, and graduated<br />

in 1995. He has since then worked<br />

with Apia XXI, where he has been the<br />

Head of the Structures Department<br />

since 2000. At the time he is finishing<br />

his Doctoral Thesis “Stability of<br />

great high piers of bridges built by<br />

cantilever method”, at the Structural<br />

and Mechanical Department at the<br />

University of Cantabria. Montabliz<br />

Viaduct has allowed him to participate<br />

in some research works such as special<br />

studies of earthquake, wind and fire;<br />

wind tunnel tests; terrain-structure<br />

interaction studies of the foundation<br />

of piers and monitoring both static and<br />

dynamic structural behaviour during<br />

its construction.<br />

Important projects he has developed<br />

include: Navas Viaduct, Caviedes<br />

Viaduct, Viaduct over Voltoya River,<br />

New Bridge over Ebro Reservoir,<br />

Santander’s Bay Ring Footbridge,<br />

Montabliz Viaduct and New Bridge<br />

over Llobregat River. He has won<br />

the Idea Tenders of the New Bridge<br />

over Guadaira River (Sevilla) and<br />

the New Bridge over Llobregat River<br />

(Barcelona). Roberto Revilla’s professional<br />

passion has always been to<br />

design bridges taking care of aesthetics<br />

and in full harmony with the surrounding,<br />

breaking civil engineering<br />

architecture barrier.<br />

Outstanding Paper Award<br />

The Outstanding Paper Award is<br />

remitted each year to the author(s)<br />

of a paper published in the preceding<br />

year’s issues of the IABSE Journal<br />

Structural Engineering International<br />

(<strong>SEI</strong>), encouraging and rewarding contributions<br />

of the highest quality. It was<br />

first launched in 1991.<br />

The Outstanding Paper Award Committee,<br />

chaired by Professor Akira<br />

Wada, Japan, has conferred the<br />

Outstanding Paper Award to Andreas<br />

Breum Ølgaard, Jens Henrik Nielsen,<br />

and John Forbes Olesen, Denmark, for<br />

their paper:<br />

“Design of Mechanically Reinforced<br />

Glass Beams: Modelling and<br />

Experiments” Published in Structural<br />

Engineering International (<strong>SEI</strong>) May<br />

2009.<br />

The paper is a study on how to obtain<br />

a ductile behaviour of a composite<br />

transparent structural element. The<br />

structural element is constructed<br />

by gluing a steel strip to the bottom<br />

face of a float glass beam using<br />

an epoxy adhesive. The composite<br />

beam is examined by four point<br />

bending tests, and the mechanisms<br />

of the beam are discussed. Analogies<br />

to reinforced concrete beam theory<br />

are made; thus, four different design<br />

criteria, depending on the reinforcement<br />

ratio, are investigated. Analytical<br />

expressions are derived that are capable<br />

of describing the behaviour in an<br />

uncracked stage, a linear cracked stage<br />

and a yield stage. A finite element<br />

model, capable of handling the cracking<br />

of the glass by killing elements, is<br />

presented.<br />

Both analytical and numerical simulations<br />

are in fairly good agreement<br />

with the experimental observations. It<br />

appears that the reinforcement ratio<br />

is limited by the risk of anchorage<br />

failure and must be adjusted accordingly<br />

to obtain safe failure behaviour<br />

in a normal reinforced mode. Analysis<br />

of anchorage failure is made through<br />

a modified Volkersen stress analysis.<br />

Furthermore, different aspects of the<br />

design philosophy of reinforced glass<br />

beams are presented.<br />

Outstanding Structure Award<br />

The Outstanding Structure Award<br />

(OStrA) was established in 1998. It is<br />

one of the highest distinctions awarded<br />

by IABSE and recognises, in different<br />

regions of the world, some of the<br />

most remarkable, innovative, creative<br />

or otherwise stimulating structures<br />

completed within the last few years.<br />

The Outstanding Structure Award<br />

476 Panorama Structural Engineering International 4/<strong>2010</strong>


Committee is chaired by Mr. William J.<br />

Nugent, USA. In <strong>2010</strong> the Outstanding<br />

Structure Award is awarded to.<br />

The National Aquatics Center,<br />

Beijing, China,<br />

for being, “a breathtaking interlocked<br />

soap bubble architecture of ETFE pillows<br />

within a polyhedral steel space<br />

frame resulting in outstanding aesthetic<br />

harmony of form function and structure<br />

which is energy efficient and pleasing<br />

to all”.<br />

Unusually in this era of architectural<br />

form making, the Beijing National<br />

Aquatics Centre, was generated as<br />

much by engineering intent as for its<br />

beauty. It is the result of an outstanding<br />

collaboration between Arup, PTW<br />

architects and CCDI. The primary<br />

purpose of the “box of bubbles” is to<br />

trap as much solar energy as possible<br />

and use it to both heat the swimming<br />

pools and light the internal spaces. This<br />

“insulated greenhouse” saves 30% of<br />

the required heating energy and 55%<br />

of the artificial lighting. This energy<br />

saving is equivalent to cladding the<br />

whole building with solar panels. It<br />

also provides a quality of internal light<br />

and space that needs to be physically<br />

experienced to be really appreciated.<br />

The structure is based on a solution<br />

to the century old mathematical<br />

“The Water Cube”, China<br />

conundrum posed by Lord Kelvin:<br />

“What is the most efficient way<br />

of subdividing three dimensional<br />

space?” This puzzle is now thought to<br />

have been solved by Professor Weaire<br />

and Dr Phelan, whose foam is also the<br />

geometry of a perfect array of soap<br />

bubbles.<br />

The geometry of Weaire-Phelan foam<br />

provided a unique structure that may<br />

be the most earthquake resistant<br />

building in the world. The building is<br />

not a pattern applied to a box, but a<br />

box carved from a theoretically perfect<br />

and repetitious array of bubbles.<br />

The final building is satisfying on<br />

many levels: as a beautiful object;<br />

a poetic expression of bubbles and<br />

water; the physical manifestation of<br />

an abstract theoretical geometry; a<br />

through thickness pattern; complementary<br />

to its neighbour, the Bird’s<br />

Nest, and uplifting to be in. But most<br />

importantly it is entirely sustainable<br />

and achieves pure engineering<br />

objectives.<br />

Outstanding Stucture Award<br />

Finalists<br />

Starting with the <strong>2010</strong> Outstanding<br />

Structure Award, IABSE is pleased to<br />

present the three Finalists selected by<br />

the OStrA Committee.<br />

Mode Gakuen Spiral Towers, Nagoya,<br />

Japan. Its design includes three towers<br />

interwined in a spiral form, suggesting<br />

the intertwined rising energy of the students<br />

of Mode Gakuen’s three schools:<br />

its fashion school (MODE), computer<br />

and animation school (HAL), and<br />

medical school (lSEN).<br />

The building has 36 floors above<br />

ground, three basement levels, and<br />

two penthouse levels. Its height is 170<br />

meters above ground and 21 meters<br />

underground. A central core having<br />

an oval cross-sectional shape consists<br />

of three wings having fan-shaped cross<br />

sections, radially arranged next to<br />

each other. The planar configuration<br />

changes with height. Three classrooms<br />

are arranged in the respective wings<br />

around the central core, which<br />

includes stairwells and elevator shafts.<br />

Ascending higher in the building in<br />

a spiral pattern, the rooms gradually<br />

become smaller in size. Displacement<br />

of the centers of rotation of the three<br />

wings produces an external appearance<br />

of organic curves.<br />

Twelve straight columns are arranged<br />

aro und this core, and braces are connected<br />

to these columns in a mesh<br />

network, forming the thick central<br />

trunk of the tubular structure (called<br />

an “inner truss tube”). This tubular<br />

structure is highly strong and rigid<br />

with regard to horizontal and twisting<br />

forces exerted on the building by<br />

earthquakes and high winds, providing<br />

the necessary structural performance.<br />

With no braces around the outside, a<br />

transparent appearance is achieved;<br />

and minimal, thin-diameter columns<br />

provide lower rigidity for a light frame<br />

that does not bear seismic forces.<br />

Mode Gakuen Sprial Towers, Japan<br />

Structural Engineering International 4/<strong>2010</strong> Panorama 477


Sutong Bridge, China. Sutong Bridge<br />

is located in the southeast of Jiangsu<br />

Province, China, which is in the lower<br />

reaches of the Yangtze River. The<br />

visionary project was motivated by<br />

the need for a highway route crossing<br />

the Yangtze River and linking<br />

Suzhou and Nantong at the opposite<br />

banks. The total length of the Bridge<br />

project is 32,4 km, consisting of three<br />

main parts, the viaducts on both banks<br />

of the river and the central part over<br />

the water, which is about 6 km long.<br />

The central part comprises of the main<br />

cable-stayed bridge with the world<br />

record 1088 m span as main navigational<br />

channel, a continuous rigid<br />

frame bridge with a main span of 268 m<br />

as secondary navigational channel, and<br />

approach bridges.<br />

The bridge substructure was designed<br />

and constructed with the emphasis on<br />

sustainable development or environmental<br />

protection in the mother river<br />

of China, Yangtze River. This aim<br />

was achieved through selecting group<br />

pile foundation instead of caisson to<br />

alleviate the impact on the river flow,<br />

installing various scour protection to<br />

minimise the erosion in the river bed,<br />

Sutong Bridge, China<br />

and adopting partially hydrolysed<br />

polyacrylamide (PHP) system for clay<br />

mud treatment to reduce the disposal<br />

of bored pile construction. One of<br />

the most significant challenges in the<br />

construction of this super-long cable<br />

stayed bridge was geometry control.<br />

The unique complexity of Sutong<br />

Bridge required specially developed<br />

methods and procedures to control<br />

the geometry profile and safety of the<br />

bridge during the construction period.<br />

Heathrow Terminal 5A, UK. The<br />

156 m clear span roof encloses three<br />

mio. sq. ft. floor space framed in steel<br />

over three storeys. The unconventional<br />

height of the building was in<br />

response to the challenge of having<br />

to build within the constraints of two<br />

runways and the greenbelt beyond<br />

them.<br />

The T5A roof is an awe inspiring<br />

structure that arches over the terminal<br />

building. The roof carries huge compression<br />

forces which are essential to<br />

prevent the buckling of its individual<br />

parts and of the structure as a whole.<br />

One of the pioneering analysis techniques<br />

employed on this project was<br />

modal buckling analysis. This calculated<br />

the effective reduction in lateral<br />

stiffness that is caused by compression<br />

forces within the structure and<br />

used eigenvector analysis to predict<br />

the most critical possible buckling<br />

modes. The mode shape data was<br />

then processed to give sets of design<br />

forces, ensuring a consistent reserve of<br />

strength against buckling, without providing<br />

extra strength where it was not<br />

needed.<br />

Heathrow T5A, UK<br />

The central arched section of the roof<br />

needed to be assembled, clad, and<br />

pre-stressed at ground level before<br />

being lifted into position using strand<br />

jacks. This creative approach was vital<br />

to ensure that the whole operation<br />

could be carried out below the airport<br />

radar ceiling and that the risk from<br />

working at height would be reduced.<br />

This idea became an integral part of<br />

the building design and construction<br />

planning of the whole Terminal.<br />

478 Panorama Structural Engineering International 4/<strong>2010</strong>


The IABSE Foundation Anton Tedesko Medal<br />

The Anton Tedesko Medal is awarded<br />

by the IABSE Foundation for<br />

the Advancement of Structural<br />

Engineering. The Award has two components:<br />

the first is a medal awarded<br />

to the Laureate in recognition of his<br />

contribution to the advancement of<br />

structural engineering. The second<br />

part is a sum of 25’000 Swiss Francs<br />

to be used by the Laureate in order<br />

to organize and finance a study leave<br />

abroad for a young promising engineer<br />

(Fellow) outside his/her home<br />

country with prestigious engineering<br />

firms. Klaus Ostenfeld, Chair<br />

of the IABSE Foundation Council,<br />

conferred the Anton Tedesko Medal<br />

to Prof. Koichi Takanashi, at the<br />

Symposium Opening Ceremony<br />

“in recognition of his dedication to<br />

excellence in structural engineering<br />

and his role as a mentor for young<br />

engineers”<br />

Koichi Takanashi has supervised many<br />

students and directed research projects<br />

at the Universities of Tokyo, Chiba<br />

and Kougakuin. His research has been<br />

focused on plastic design and seismic<br />

design. One of his outstanding accomplishments<br />

was the establishment of<br />

the overall testing method to combine<br />

numerical analysis of structural<br />

system in the computer and the test<br />

of structural frames in 1974. This<br />

method has advanced research to precisely<br />

understand the behaviour of<br />

structural frames. This method is now<br />

widely used in the world and developed<br />

as one of the standard methods<br />

in earthquake response analysis<br />

and structural testing. As his important<br />

role in the structural engineering<br />

society, he chaired the ‘Structural<br />

Design Appraisal Committee of Tall<br />

Buildings’ for eight years, where structural<br />

design works of all tall buildings<br />

in Japan were examined and appraised<br />

from the viewpoint of the structural<br />

performance against earthquake.<br />

Also, he directs big research projects.<br />

His latest project is “Development of<br />

a new structural system” which was<br />

reported in <strong>SEI</strong> Vol. 20 No. 1. Prof.<br />

Takanashi is currently President of<br />

the Japan Society of Structural Steel<br />

Construction (JSSC), Member of the<br />

Architectural Institute of Japan (AIJ)<br />

and the International Association for<br />

Bridge and Structural Engineering<br />

(IABSE).<br />

Koichi Takanashi, Japan<br />

Within IABSE Koichi Takanashi has<br />

contributed comprehensively since<br />

his first presentation of his paper in<br />

IABSE Symposium Lisbon in 1973.<br />

He has submitted numerous papers<br />

and given his support to an IABSE<br />

Congress and Symposium as an Invited<br />

and a Keynote Speaker. He participated<br />

in Working Commission 8 and<br />

5 as a member. He convened IABSE<br />

Symposium Davos, Rome and Kobe as<br />

a Member of the Scientific Committee.<br />

He further has extended his efforts<br />

as a Vice-President from 1997–2005,<br />

served as Chair of the Japanese Group<br />

from 1999–2005, he was also a Member<br />

of the IABSE Outstanding Structure<br />

Award Committee.<br />

IABSE Symposium Venice, September 22–24, <strong>2010</strong><br />

Large Structures and Infrastructures for Environmentally Constrained and Urbanised Areas<br />

Venice <strong>2010</strong> Symposium Banner<br />

The Italian Group of IABSE, chaired<br />

by Carlo Urbano, welcomed the<br />

world’s structural engineers and their<br />

accompanying persons to an important<br />

event that promoted science<br />

and practice in Bridge and Structural<br />

Engineering at its highest levels. For<br />

the IABSE Symposium it was a return<br />

to Venice, after it had been hosted in<br />

the Serenissima for the first time back<br />

in 1983.<br />

The Organising Committee was chaired<br />

by Enzo Siviero, Italy, the Scientific<br />

Committee by Massimo Majowiecki,<br />

Italy and the Advisory Committee by<br />

Anton Steffen, Switzerland. The excellent<br />

Secretaries were Bruno Briseghella<br />

for the Organising and Tobia Zordan<br />

for the Scientific Committee. The symposium<br />

topic had found the interest<br />

of a considerable number of Italian<br />

Universities, including the Istituto<br />

Structural Engineering International 4/<strong>2010</strong> Panorama 479


and on a CD. The book (899<br />

pages) and CD can be ordered at:<br />

www.iabse.org/publications/onlineshop/<br />

BASAAR<br />

Several IABSE Working Commissions’<br />

and Working Groups held<br />

their annual ‘BASAAR’ (Briefings<br />

About Structural Applications and<br />

Research) on Wednesday afternoon,<br />

September 22. The purpose of the<br />

BASAAR is to promote a lively discussion<br />

about a topic predetermined<br />

by each Working Commission or<br />

Group. Following topics were presented<br />

and discussed:<br />

Carlo Urbano, Chair, Italian Group of IABSE<br />

Universitario di Architettura di<br />

Venezia and the Polictecnico di Milano<br />

as co-organisers, and the Universities<br />

of Naples, Rome, Trento and Turin as<br />

supporters.<br />

The Palazzo del Casino and the<br />

Palazzo del Cinema, at the Lido in<br />

Venice served as venues. The attendance<br />

was high with almost 600 participants<br />

including accompanying<br />

persons. Out of 400 abstracts submitted<br />

to the theme ‘Large Structures and<br />

Infrastructures for Environmentally<br />

Constrained and Urbanised Areas’<br />

the Scientific Committee selected 210<br />

papers for oral presentation and some<br />

150 for poster presentation, from 47<br />

countries.<br />

The Opening Ceremony<br />

Welcome and Introductory Speeches<br />

were made by Carlo Urbano, Chair<br />

of the Italian Group of IABSE; Enzo<br />

Siviero, Chair of the Organising<br />

Committee; Antonio Paruzzolo representing<br />

Giorgio Orsoni, Mayor of<br />

Venice, Don Alberto representing<br />

Angelo Scola, Cardinal of Venice, and<br />

Jacques Combault, President IABSE,<br />

who subsequently presented the<br />

IABSE Awards <strong>2010</strong>. Klaus Ostenfeld,<br />

Chair of the IABSE Foundation<br />

Council then conferred the IABSE<br />

Anton Tedesko Medal. Alberto<br />

Scotti gave an invited lecture on ‘The<br />

Venice Mose Project: an Holistic<br />

and Interdisciplinary Approach for<br />

Innovative Interventions’ and completed<br />

the Ceremony.<br />

Symposium Contents<br />

Keynote Lectures, Presentations and<br />

Posters addressed the following main<br />

topics: Basis of Design; Infrastructure<br />

and Design as Meeting Point for<br />

Architecture and Engineering; Infrastructure<br />

Hazard and Safety Concepts;<br />

Management and Planning of Operation<br />

and Maintenance; Ethics and<br />

Social Responsibility.<br />

Keynote Lectures<br />

• Giorgio Diana, Italy<br />

The Messina Strait Bridge: Major<br />

Problems Affecting the Design<br />

• Klaus Ostenfeld, Denmark<br />

An Integrated Multidisciplinary Approach<br />

to Design of Major Fixed Links<br />

• Jiemin Ding, China<br />

Recent Applications and Practices of<br />

Large-Span Steel Structures in China.<br />

Invited Lectures were given on the<br />

following days of the Symposium:<br />

‘Traceability Systems and Quality<br />

Systems: two Sides of same Coin’, by<br />

Corrado Baldi; ‘The Raising of the<br />

Buildings of the City of Venice for the<br />

Safeguard from the High Water’, by<br />

Enzo Siviero.<br />

Symposium Report<br />

Three Keynote Lectures and<br />

375 contributions have been collected<br />

in the Symposium Report<br />

The Long-term Structural<br />

Performance (Life-Cycle Costs)<br />

by F. Biondini, Italy and M. Torkkelli<br />

Finland,<br />

(Co-ordinator M. SØderkvist, Finland,<br />

WC1)<br />

Monitoring – Does it make Sense?<br />

by R. Geier, Austria and S. Nakamura,<br />

Japan,<br />

(Co-ordinator: R. Geier, Austria, WC 2).<br />

Structural Safety Assessment of<br />

Concrete Structures<br />

by J. McGormley, USA; M.Matsumoto<br />

Japan and C. Bob, Romania,<br />

(Co-ordinator: J. Tortorella, USA, WC 4).<br />

The Long Term Issues Facing<br />

Structural Engineers in<br />

Environmentally Constrained Urban<br />

Areas<br />

by A. Boegle, Germany; A. Meyboom,<br />

Canada and F. Saad, Egypt,<br />

(Co-ordinator: W. Anderson, Canada,<br />

WC 5)<br />

Integrating Sustainability into<br />

Structures<br />

by J. Kanda, X. Ruan, China and<br />

J. Anderson, USA,<br />

(Co-ordinator: J. Kanda, Japan, WC 7)<br />

Seismic Resistance of Structures –<br />

Lessons from Devastating<br />

Earthquakes (Chile, Mexico, Turkey,<br />

Haiti, Greece)<br />

by L.F. Fargier Gabaldon, USA;<br />

C. Mendez, Switzerland; M. Gercek,<br />

Turkey and D. Sonda, Italy,<br />

(Co-ordinator: St. Dritsos, WG 7)<br />

In addition to the BASAAR a video<br />

presentation ‘Vibrations of the Volga<br />

Bridge in Volgograd, Russia’ was<br />

organised by S. Mozalev, Chair of the<br />

Russian Group of IABSE.<br />

480 Panorama Structural Engineering International 4/<strong>2010</strong>


Young Engineers Programme<br />

Benefits for young engineers at this<br />

Symposium were jointly offered by<br />

the IABSE Fellows and the Symposium<br />

Organising Committee:<br />

Young Engineers enjoyed reduced<br />

registration fee and were offered<br />

free IABSE membership for the<br />

year 2011. The Award Jury represented<br />

by A. Chen, China, P. Collin,<br />

Sweden and R. Zandonini, Italy, conferred<br />

two prizes each of 2000 EUR,<br />

sponsored by IABSE Fellows to two<br />

young authors for their outstanding<br />

contributions:<br />

Johan Berger, Austria: ‘New Ap proach<br />

for Bridges with Very High Durability’<br />

and Xin Ruan, China: ‘Failure Analysis<br />

of a Long Span Pre-stressed Concrete<br />

Box Girder Bridge’.<br />

YEP Awardees Johan Berger, Austria<br />

and Xin Ruan, China<br />

Technical Visits<br />

On Friday a technical visit was organised<br />

to the moveable dams of ‘MOSE’,<br />

a defence system consisting of rows of<br />

concealed gates designed to stop high<br />

tides at the three lagoon inlets. On<br />

Saturday another technical excursion<br />

to three structures: the fourth bridge<br />

over the Grand Canal in Venice,<br />

designed by Santiago di Calatrava; the<br />

Ponte Strallato di Marghera, a curved<br />

cable-stayed bridge linking the city of<br />

Mestre to Porto Marghera and to the<br />

‘Laguna Palace’ glass roof: a complex<br />

constituted by a hotel building and a<br />

private dock.<br />

Special Public Session<br />

On Friday, September 24, after the<br />

Symposium a Special Session on<br />

‘Recent Large Earthquake and Seismic<br />

Risk Reduction with Rreference to<br />

a Sustainable Development’ was<br />

organised, and open for free to the<br />

Public.<br />

Social Events and Sightseeing<br />

On Wednesday, the Organising Committee<br />

welcomed all participants and<br />

accompanying persons to the beautiful<br />

Chiostro di San Nicoletto, dating to<br />

the origins of the independent Venice,<br />

in the early Middle Ages. Drinks and<br />

food and a wonderful soprano recital<br />

were enjoyed in a white decoration<br />

at candle-light. Later in the evening<br />

unexpected visitors interrupted the<br />

peaceful atmosphere and initiated<br />

what was later called the ‘mosquito’<br />

dance, which many guests abandoned<br />

to continue with a nice dinner in<br />

Venice or to join the young engineers<br />

social event at the historical Nicelli<br />

Lido Airport for free drinks and music<br />

entertainment.<br />

Prior to the Symposium Dinner on<br />

Thursday evening, a visit to either<br />

San Marco’s Church with a magnificent<br />

organ concert or to the Scuola<br />

Grande di San Rocco were organised.<br />

The evening was continued with a<br />

fine six course menu at the beautiful<br />

Cà Giustinian Palace, the seat of the<br />

Biennale, located close to San Marco<br />

Square and overlooking the Grand<br />

Canal.<br />

Welcome Reception at Chiostro di San Nicoletto<br />

Social Tours<br />

Symposium delegates and accompanying<br />

persons were offered various<br />

guided tours in Venice by foot or<br />

Gondola, outings to Murano, Burano,<br />

or longer excursions to cities such<br />

as Padova, Verona, Florence and<br />

Rome.<br />

Members of the Organising Committee and Guests<br />

Structural Engineering International 4/<strong>2010</strong> Panorama 481


Dhaka Conference ‘Advances in Bridge Engineering-II’<br />

Organised by the Bangladesh Group of IABSE and JSCE Steel Structures Committee, Japan<br />

The Dhaka Conference on ‘Advances<br />

in Bridge Engineering-II’ held from<br />

August 8–10, <strong>2010</strong>, was a great success.<br />

It was jointly organised by the<br />

new Bangladesh Group of IABSE<br />

and JSCE Steel Structures Committee,<br />

in Association with the Roads and<br />

Highways Department Government<br />

of Bangladesh and the Institution<br />

of Engineers, Bangladesh and<br />

co-organisers from the Bangladesh<br />

Sections of the American Society<br />

of Civil Engineers and Institution<br />

of Civil Engineers, UK. One of the<br />

major aspects of this conference<br />

was bringing major professional<br />

societies and bodies working in the<br />

region together to achieve an effective<br />

collaboration and partnership in<br />

future.<br />

B.C. Roy, Vice-President of IABSE<br />

presented a message on behalf of<br />

IABSE President, Jacques Combault<br />

at the Opening Ceremony, and two<br />

Honorable Ministers responsible for<br />

the Ministries of Communication,<br />

Science and Information Technology,<br />

Government of Bangladesh, attended<br />

the Conference Dinner and Closing<br />

Ceremonies of the conference. The<br />

presence of Muhammad Yunus, Nobel<br />

Laureate for Peace in the Conference<br />

Dinner inspired participants greatly.<br />

Experts, professionals and academicians<br />

from Asia, Europe, Australia<br />

and the North America discussed the<br />

issues of Bridge Engineering and its<br />

further development for the benefit of<br />

the people and community. The total<br />

number of registration to the conference<br />

was 338.<br />

Six Keynote addresses were supported<br />

by 56 technical papers from<br />

11 countries and four continents. Bridge<br />

engineering themes were on: History<br />

and Planning Materials; Analysis;<br />

Design and Construction Design for<br />

Dynamic Forces; Geo-environmental<br />

Issues and Administration; Maintenance<br />

and Monitoring. Ref.: www.<br />

iabse-bd.org<br />

Conference participants perceived the<br />

technological challenges that exist in<br />

Bangladesh for bridge construction and<br />

maintenance, but also recognised the<br />

necessity of having world class bridges<br />

in Bangladesh for the effective transportation<br />

not only within the country,<br />

but also within the Asian region. They<br />

identified the necessity of technology<br />

transfer, assimilation and formulation<br />

of a unified bridge code for the Asian<br />

countries in the coming years.<br />

A volume with 594 pages has been<br />

published, for more information,<br />

please contact A.F.M. Saiful Amin:<br />

samin@ce.buet.ac.bd<br />

Md. Yunus (right), Nobel Laureate for Peace. From left:<br />

H. Mutsuyoshi, Saitama Univ.; M. Nagai, Nagoaka Univ.<br />

of Tech. Japan; K. Wheeler, Maunsell AECOM<br />

Left to right: M. A. Sobhan (Chair, Bangladesh IABSE Group);<br />

Md. Yunus (Nobel Laureate); A.F.M. Saiful Amin (General<br />

Secretary, Bangladesh IABSE Group)<br />

482 Panorama Structural Engineering International 4/<strong>2010</strong>


Maschinenfabrik GmbH & Co. KG<br />

A matter of confidence.<br />

Stressing of stay cables<br />

PAUL supplies<br />

• Prestressing installations incl. planning work<br />

• Anchor grips<br />

• Prestressing machinery<br />

(single-/multi-stressing jacks)<br />

• Strand pushing and cutting equipment<br />

• Automatic prestressing machines for railway sleepers<br />

• Prestressing equipment for bridge construction<br />

(prestressing cables and stay cables)<br />

The experts in<br />

Prestressed Concrete Technology.<br />

Max-Paul-Straße 1<br />

88525 Dürmentingen / Germany<br />

www.paul.eu<br />

Phone: +49 (0) 7371 / 500 - 0<br />

Fax: +49 (0) 7371 / 500 - 111<br />

Mail: stressing@paul.eu


Brilliant Slide<br />

MSA ® – New for<br />

spherical bearings!<br />

MAURER Sliding Alloy<br />

New alloy MSA ® for bearings in<br />

bridge and building structures.<br />

Shining – special treatment of<br />

the surface for reduced sliding<br />

resistance and abrasive wear<br />

Resistant to agressive<br />

industrial air – especially against<br />

fluorine and chlorine<br />

Superior corrosion resistance<br />

Better economic efficiency<br />

Longer lifetime<br />

Approved by the construction<br />

authorities (AbZ Z-16.4-436)<br />

Our Product Range<br />

Maurer Expansion Joints<br />

Maurer Structural Bearings<br />

Maurer Seismic Protection Systems<br />

Maurer Vibration Control Systems<br />

Maurer Monitoring of Structures<br />

Maurer Söhne GmbH & Co. KG<br />

Frankfurter Ring 193, 80807 Munich/Germany<br />

Phone +49 89 32394–0<br />

Fax +49 89 32394–306<br />

ba@maurer-soehne.de<br />

www.maurer-soehne.com

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!