You are on page 1of 118

Concrete Society Technical Report No.

55
Second Edition

Design guidance for


strengthening concrete
structures using fibre
composite materials
Second Edition

World Leaders in
Composite Solutions

High Strength
Lightweight
Non-corrosive
Easy Installation
Flexible

A comprehensive approach to single source support from specification through to completion.


For further information please contact Paul Russell on 0161 794 7411 or email
paul.russell@degussa.com
Degussa Construction Chemicals (UK),
Albany House, Swinton Hall Road, Swinton, Manchester M27 4DT
Tel: 0161 794 7411 Fax: 0161 727 8881
e-mail: mbtfeb@degussa.com www.degussa-cc.co.uk

Concrete Society Technical Report No. 55


Second Edition

Design guidance for


strengthening concrete
structures using fibre
composite materials
Second Edition

Report of a Concrete Society Committee

The Concrete Society

Design guidance for strengthening concrete structures using fibre composite materials
Concrete Society Technical Report No. 55
ISBN 1 904482 14 7
The Concrete Society 2004
First edition published 2000
Further copies and information about membership of The Concrete Society may be obtained from:
The Concrete Society
Riverside House, 4 Meadows Business Park
Station Approach, Blackwater
Camberley, Surrey GU17 9AB, UK
E-mail: enquiries@concrete.org.uk; www.concrete.org.uk
All rights reserved. Except as permitted under current legislation no part of this work may be photocopied, stored
in a retrieval system, published, performed in public, adapted, broadcast, transmitted, recorded or reproduced in
any form or by any means, without the prior permission of the copyright owner. Enquiries should be addressed
to The Concrete Society.
The recommendations contained herein are intended only as a general guide and, before being used in connection
with any report or specification, they should be reviewed with regard to the full circumstances of such use.
Although every care has been taken in the preparation of this Report, no liability for negligence or otherwise can
be accepted by The Concrete Society, the members of its working parties, its servants or agents.
Concrete Society publications are subject to revision from time to time and readers should ensure that they are in
possession of the latest version.
Printed by Cromwell Press, Trowbridge, Wiltshire

CONTENTS

Members of the Project Committee


List of Figures
List of Tables
Executive Summary

viii
x
x
xi

INTRODUCTION

BACKGROUND

2.1
2.2
2.3
2.4

3
3
4

2.7

Principles of strengthening
Assessment of structures to be strengthened
Strengthening solutions
Advantages and disadvantages of fibre composite
strengthening
2.4.1 Advantages
2.4.2 Disadvantages
Design life
Economics
2.6.1 Installation
2.6.2 Whole-life costing
Level of strengthening

5
5
6
7
7
7
8
8

MATERIAL TYPES AND PROPERTIES

Introduction
Fibres
3.2.1 Types of fibre
3.2.2 Performance of different types of fibre
3.3 Fabrics
3.4 Plates
3.5 Rods and strips for near-surface-mounted (NSM)
reinforcement
3.6 Preformed shells for column confinement
3.7 Specials
3.8 Adhesives and laminating resins
3.9 Environmental aspects and Health and Safety
3.9.1 Environmental aspects
3.9.2 Health and Safety
3.10 Choice of materials for design
3.10.1 Plates versus wet lay-up sheet systems
3.10.2 NSM systems
3.10.3 Specific composite material
3.10.4 Stiffness issues

9
9
9
9
10
11

2.5
2.6

3.1
3.2

REVIEW OF APPLICATIONS

15

4.1
4.2

Introduction
Buildings
4.2.1 Beams and slabs
4.2.2 Columns
4.2.3 Connections
4.2.4 Walls
Bridges
4.3.1 Beams and slabs
4.3.2 Columns
4.3.3 Continuity
Other structures
4.4.1 Towers and chimneys
4.4.2 Tunnels
4.4.3 Marine/coastal structures
4.4.4 Miscellaneous structures

15
16
16
17
18
18
18
18
20
21
21
21
21
21
22

STRUCTURAL DESIGN OF
STRENGTHENED MEMBERS

23

5.1
5.2
5.3
5.4

Symbols
Overview of available design guidance
Basis of design
Mechanical properties of materials
5.4.1 Properties of concrete and steel
reinforcement
5.4.2 Properties of fibre-reinforced
polymers (FRP)
5.4.3 Properties of adhesives and laminating
resins
5.4.4 Stressstrain curves
Partial safety factors for loads
Design values for material properties
5.6.1 Introduction
5.6.2 Design strength of steel and concrete
5.6.3 Design elastic modulus of FRP
5.6.4 Design ultimate strain of FRP
5.6.5 Design ultimate strength of FRP
5.6.6 Steel stress
5.6.7 Deflection and cracking
5.6.8 Adhesive

4.3

4.4

11
11
12
12
13
13
13
13
13
13
14
14

5.5
5.6

23
24
24
25
25
25
26
26
26
26
26
26
26
27
27
27
28
28

Extreme loadings
5.7.1 Behaviour of structures in fire
5.7.2 Seismic loading
5.7.3 Impact loading
5.7.4 Blast loading
5.7.5 Vandalism

28
28
28
29
29
29

SHEAR STRENGTHENING

47

7.1
7.2

Introduction
Design procedure
7.2.1 Maximum shear capacity
7.2.2 FRP shear strengthening design
Spacing of FRP strips
Additional axial FRP

47
47
47
48
49
49

STRENGTHENING MEMBERS IN
FLEXURE

31

6.1
6.2

General
Moment capacity
6.2.1 Introduction
6.2.2 Requirements of the existing section
6.2.3 Preliminary design
6.2.4 Design resistance moment of FRPstrengthened beam
6.2.5 Example design method
FRP separation failure
6.3.1 Introduction
6.3.2 Bond failure
6.3.3 Design procedure
Flexural strengthening with NSM reinforcement
6.4.1 Introduction
6.4.2 Design basis
6.4.3 Bond behaviour
6.4.4 Modes of failure
6.4.5 NSM separation failure design
6.4.6 Anchorage design
Flexural strengthening plate location
Thick and multi-layer laminates
Redistribution
Serviceability
6.8.1 Crack widths
6.8.2 Deflections and material stresses
6.8.3 Fatigue
6.8.4 Stress rupture
6.8.5 Strengthening under non-static live load
Strengthening prestressed structures

31
31
31
31
32

STRENGTHENING AXIALLY LOADED


MEMBERS
53

8.1
8.2
8.3

Introduction
Stressstrain model for FRP-confined concrete
Compression
8.3.1 Introduction
8.3.2 Tensile rupture of FRP
8.3.3 Lap joint failure
8.3.4 Shear
8.3.5 Serviceability
Flexure
8.4.1 Introduction
8.4.2 Moment capacity with axial load
8.4.3 Debonding
8.4.4 Anchorage
Ductility
Strengthening columns with non-circular crosssection

53
53
55
55
55
56
56
57
57
57
58
59
59
59

EMERGING TECHNOLOGIES

65

9.1

Emerging strengthening technologies already


used in practice
9.1.1 Prestressing using FRP composites
9.1.2 Unstressed FRP anchorage techniques
9.1.3 Bolted plate anchors
9.1.4 Concrete masonry walls
Emerging strengthening technologies at the
research stage
9.2.1 Prestressed NSM bars
9.2.2 NSM bars for shear strengthening

5.7

6.3

6.4

6.5
6.6
6.7
6.8

6.9

vi

32
32
33
33
34
34
36
36
37
37
37
38
38
39
39
40
40
40
40
41
41
41
42

7.3
7.4

8.4

8.5
8.6

9.2

60

65
65
67
67
68
68
68
68

9.2.3
9.2.4
9.2.5

FRP anchor systems


Steel-reinforced polymers
Deep embedded bar for shear
strengthening
9.2.6 Prestressed carbon FRP straps for shear
strengthening
9.2.7 Mechanical fastening techniques
9.2.8 Strengthening for torsion
9.2.9 Life expectancy modelling
9.2.10 Inorganic adhesives

10

68
68
69
69
69
69
69
69

10.8 Assembly and visual inspection


10.8.1 Installation of FRP plates
10.8.2 Installation of FRP fabrics
10.8.3 Installation of NSM reinforcement
10.9 Control samples
10.10 Non-destructive tests
10.11 Application of over-coatings
10.12 Identification/warning signs
10.13 Records

11

LONG-TERM INSPECTION, MONITORING


AND MAINTENANCE
79

11.1
11.2
11.3
11.4
11.5

Inspection and monitoring regime


Frequency of inspections
Routine visual inspection
Detailed inspection
Maintenance

WORKMANSHIP AND INSTALLATION 71

10.1 Introduction
10.2 Evaluation of concrete condition
10.3 Concrete preparation
10.3.1 Concrete surface for plates and fabric
10.3.2 Slots in concrete surface for NSM
material
10.4 Material conformity
10.5 Storage of materials
10.6 Site conditions
10.7 Mixing and application of adhesive
10.7.1 General
10.7.2 Application to substrate prior to plate
installation
10.7.3 Application to FRP plates
10.7.4 Application to substrate prior to fabric
installation
10.7.5 Application to FRP fabrics
10.7.6 Inserting adhesive into slots for NSM
reinforcement

71
71
72
72
73
73
73
73
74
74
74
74
75
75

75
75
76
77
77
77
77
78
78

79
79
80
80
80

REFERENCES

81

APPENDICES

89

A
B
C
D

89
91
95

Glossary of terms
Systems available in the UK
Quality control of materials
Specialist suppliers, contractors, consultants,
universities/research organisations and owners

INDEX

97

99

75

vii

MEMBERS OF THE PROJECT COMMITTEE


Neil Loudon
Brian Bell
Bob Berry
Peter Brown
Lee Canning
John Clarke
Antony Darby
Steve Denton
John Drewett
Edward Donnelly
Neil Farmer
Alan Hardie
Chris Hughes
Tim Ibell
Sarah Kaethner
John Keble
Mike Langdon
Sam Luke
Mick Mahon
Peter Milligan
Jim Moriarty
Steve Richards
Martin Richardson
Paul Russell
Jon Shave
Ian Smith
Simon Walters

Highways Agency (Chairman)


Network Rail
Concrete Repairs Limited (representing the Concrete Repair Association)
Oxfordshire County Council (representing the CSS)
Mouchel Parkman
The Concrete Society (Secretary)
University of Bath
Parsons Brinckerhoff
Concrete Repairs Limited (representing the Concrete Repair Association)
Fyfe Co. LLC
Tony Gee and Partners
Network Rail
weber building solutions
University of Bath
Arup R & D
weber building solutions
Degussa Construction Chemicals
Mouchel Parkman
Toray Europe Ltd
Fyfe Co. LLC
London Underground
Exchem Mining and Construction
Sika
Degussa Construction Chemicals
Parsons Brinckerhoff
Tony Gee and Partners
Mouchel Parkman

CORRESPONDING MEMBERS
Michael Johnston
Tony McNulty
Wendel Sebastian
David Tann

viii

British Energy Generation UK Ltd


Health & Safety Executive Nuclear Inspectorate Directorate
University of Bristol
University of Glamorgan

ACKNOWLEDGEMENTS
The work of preparing this Report was funded by the following organisations:
Degussa Construction Chemicals
Exchem Mining and Construction
Fyfe Co. LLC and Fyfe Asia Pte.
Highways Agency
London Underground Ltd
Network Rail
Sika
Toray Europe
weber building solutions
The Concrete Society is grateful to the following for providing photographs for inclusion in the Report:
Concrete Repairs Ltd (Figures 5, 6, 8, 11, 44 and 52)
Cornwall County Council (Figures 17, 37, 39, 45 and 50)
Halcrow Group Ltd (Figures 3, 40 and 43)
Highways Agency (Figure 9)
Makers UK Ltd (Figures 4, 13, 38 and 42)
Maunsell structural Plastics (Figures 12 and 48)
Parsons Brinckerhoff (Figure 15)
Sika Ltd (Figures 7, 14, 46 and 49)
weber building solutions (Figures 10, 41 and 47)

ix

LIST OF FIGURES
Figure 1: Different types of structural strengthening,
applied to beams, slabs, walls and columns.
Figure 2: Flow chart of assessment process.
Figure 3: Installing fibre composite plates in a culvert.
Figure 4: Installing FRP plate, showing the flexibility of
the material.
Figure 5: FRP plate installed behind existing services.
Figure 6: Overlapped carbon FRP plates on Dudley Port
Bridge, West Midlands.
Figure 7: Coil of carbon FRP plate.
Figure 8: Cutting carbon FRP plate on site.
Figure 9: Shear reinforcement straps.
Figure 10: Strengthening around hole cut through slab.
Figure 11: Underpass, Great Missenden.
Figure 12: Applying carbon fibre sheet to Greenbridge
Subway, Swindon.
Figure 13: Strengthening Glade Bridge.
Figure 14: Strengthening the top surface of a bridge using
carbon fibre plates.
Figure 15: Column wrapping.
Figure 16: Machine for wrapping columns.
Figure 17: Bible Christian Bridge, Cornwall.
Figure 18: Assumed stressstrain curves.
Figure 19: Strengthening beams and slabs with FRP.
Figure 20: Stressstrain curve for reinforcing steel in the
design of strengthened beams in flexure.
Figure 21: Possible failure modes and locations for FRPstrengthened beam.
Figure 22: Variation in FRP separation strain with bonded
length, based on Denton et al.(112).
Figure 23: Characteristic bond failure force vs anchorage
length.
Figure 24: Shear reinforcement configurations.
Figure 25: General notation for shear strengthening.
Figure 26: Typical variation in ultimate strain capacity with
bonded length (after Neubauer and Rostasy(111)).
Figure 27: Experimental verification of design method
(after Denton et al.(112)).
Figure 28: Idealised stressstrain curve for FRP-confined
concrete.
Figure 29: Stressstrain model.
Figure 30: Laps in columns.
Figure 31: Stressstrain behaviour of variably confined
concrete.
Figure 32: Recommended stressstrain curves for combined flexure and axial load.
Figure 33: Assumed confined region for FRP-wrapped rectangular column.
Figure 34: Overlapping parabolas in confined region.
Figure 35: Dead-end prestressing system (left) and stressing anchorage (right).
Figure 36: Prestressing system used on Lauterbridge,
Gomadingen.
Figure 37: Use of pull-out test to determine concrete
strength.

Figure 38: Surface grinding.


Figure 39: Filling imperfections with quick-setting repair
mortar.
Figure 40: Pull-off specimen after removal from concrete
surface.
Figure 41: Mixing adhesive.
Figure 42: Application of adhesive to concrete surface.
Figure 43: Application of adhesive layer on to fibre composite plate.
Figure 44: Cutting fabric.
Figure 45: Applying resin using roller.
Figure 46: Impregnation of fabric.
Figure 47: Installing FRP plates, using a roller to apply
pressure.
Figure 48: Wrapping fabric round an arched member.
Figure 49: Wrapping fabric round column.
Figure 50: Rolling fabric to consolidate layers.
Figure 51: Double-lap shear test.
Figure 52: Spray application of mortar over-coating.
Figure 53: Example of warning printed on carbon fibre
plate.
Figure 54: Examples of proposed warning plates fixed to
structure adjacent to strengthened area.

LIST OF TABLES
Table 1:
Table 2:
Table 3:
Table 4:
Table 5:
Table 6:

Table 7:
Table 8:
Table 9:

Table 10:
Table B1:
Table B2:
Table B3:
Table B4:
Table B5:
Table B6:

Typical dry fibre properties.


Examples of strengthening of buildings in the
UK.
Examples of strengthening of bridges in the UK.
Limit states relevant to FRP strengthening systems.
Partial safety factors for Youngs modulus at the
ultimate limit state.
Recommended values of additional partial safety factors, to be applied to manufactured composites, based on Clarke(94).
Partial safety factor for strain at the ultimate
limit state.
Maximum stress ranges as a proportion of the
design ultimate strength (%).
Maximum stress under service loads to avoid
stress rupture as a proportion of design strength
(%).
Reduction in strength of adhesive for given liveload strains at FRPconcrete interface (%).
Suppliers of strengthening materials.
Properties of fibre composite sheet materials.
Properties of composite plate materials.
Properties of NSM rods and strips.
Properties of epoxy adhesives.
Properties of laminating resins.

EXECUTIVE SUMMARY
Fibre composites (or fibre-reinforced polymers, generally known as FRPs) have been used successfully for many years in the
aerospace and automotive industries. They are used in construction, for example as structural elements and for cladding. This
Report does not consider such applications but deals only with a recent development, strengthening concrete structures by
bonding fibre composites to the surface.
Suitable fibres are made from carbon, aramids or glass. These may be used in the form of:

composite plates, made from fibres and epoxy resins which are fixed with epoxies to the soffits of beams and slabs
sheet materials, which are wrapped round columns and similar members
preformed shells, bonded round columns.

Advantages
The principal advantages of using composites over steel plates are their high strength and light weight; typical properties are
given for commercially available materials. This makes installation simple and quick and eliminates the need for temporary
support. The materials can be easily cut to length on site. The availability of long lengths and the flexibility of the materials
also simplify installation because:

laps and joints are not required


the material can take up irregularities in the shape of the concrete surface and can follow a curved profile
the material can be readily installed behind existing services
overlapping, required when strengthening in two directions, is not a problem because the material is thin.

These various factors in combination lead to a significantly simpler and quicker strengthening process than when using other
methods. This is particularly important for bridges because of the high costs of lane closures and possession times on major
highways and railway lines. An additional advantage of FRPs over some other types of strengthening is that the weight of the
structure and the dimensions of the member are not significantly increased. The latter may be particularly important for
bridges, tunnels and other structures with limited clearance.
Disadvantages
One disadvantage of FRP strengthening is the risk of fire, vandalism or accidental damage. For bridges over roads the risk of
soffit reinforcement being hit by over-height vehicles should be considered. In general, some form of protection will be
required.
Examples of FRP strengthening
There are many concrete structures around the world that have been externally strengthened with FRP. The Report concentrates on applications in the UK. The floors of various buildings have been strengthened to carry additional loads and FRP
has been used in structural alterations. Columns have been strengthened in several multi-storey car-parks by wrapping with
carbon fibre sheet.
Several major highway bridges and a large number of small bridges have been strengthened using FRPs to increase their load
capacity. Most applications have been on soffits but some bridges have had FRP bonded to the upper surface or around the
columns. Other strengthening applications in the UK include lighthouses and cooling towers. Elsewhere in the world almost
every type of concrete structure, from chimneys to tunnels, has been strengthened.
Design approach
Fibre composites have a straight-line stressstrain response to ultimate with no yielding. Thus elastic methods of analysis
with no redistribution are appropriate. For members in bending, the traditional design assumptions are still valid. However,
further checks are required to avoid peeling failure at the ends of the laminate and debonding from the concrete. If failure
occurs, it will be in the outer layer of the concrete; the proposed, conservative, approach is to limit the longitudinal shear
stress in the concrete at ultimate to 0.8N/mm2. To minimise the risk of debonding, the strain in the FRP should not exceed
0.8% when the applied load is uniformly distributed and 0.6% if combined high shear forces and bending moments are present. A minimum anchorage length of 500mm is recommended.

xi

FRP strips may be used to strengthen members in shear. The material may be treated as an external stirrup, again using traditional design assumptions but the strain in the FRP should be limited to 0.4%.
Wrapping circular columns with FRP increases the axial load capacity as well as the bending and shear capacities. (Only limited increases are possible with square and rectangular columns.) Approaches are given which relate the enhanced ultimate
stress and strain in the concrete to the degree of confinement.
Workmanship and installation
The installation of FRP materials must carried out correctly, to ensure good long-term performance. Detailed guidance is
given, including the selection of the appropriate material and adhesive, adequate preparation of the concrete surface, application of the composite and correct curing of the adhesive. It is important that the work is carried out by a suitably qualified
contractor with suitably trained staff.
Inspection and maintenance
As strengthening with FRPs is a relatively new technique, regular inspection and maintenance regimes should be set up. This
is particularly important for buildings which, unlike bridges, are not generally subject to any form of routine inspection.
Where practical, additional material should be installed, which can be removed at a later stage for testing. Information on the
materials used, along with information on the actions to be taken in the event of damage to the FRP, should be included in the
Health and Safety File.
Changes and additions in the Second Edition of TR55
Since the publication of the First Edition of TR55 in December 2000 materials and techniques have developed rapidly, along
with the range of applications. Hence it was thought necessary to produce this Second Edition of TR55. A number of the
changes are matters of detail, brought about by additional research findings and further experience of the use of the materials.
However, significant changes or additions have been made in some areas, including:

modification of the treatment of partial safety factors in the design process; factors are now applied to the FRP strains rather
than the stresses
extreme loadings
design of members in shear, to provide a less empirical approach, allowing wider and more confident application of the
technique
column design, to provide a more unified approach for axial and combined axial and flexural strengthening; a more detailed
approach has been developed for the strengthening of rectangular columns
design guidance for the new technique of near-surface-mounted (NSM) reinforcement.

In addition, an overview of emerging technologies, such as the use of prestressed composites, mechanical anchorage systems
and alternatives to the adhesives currently used, has been included.

xii

INTRODUCTION

Fibre composites have been successfully used for many


years in the aerospace and automotive industries. They are
also used in construction, for example, for structural
elements, particularly in aggressive environments such as
chemical plants, and for cladding. This Report does not
consider such applications but deals only with a relatively
recent development, the strengthening of concrete structures
with fibre composite materials bonded on or near the
surface.
There are a number of situations where the load-carrying
capacity of a structure in service may need to be increased,
such as change of loading or use, or where the structure has
been damaged. In the past, strength would be increased by
casting additional reinforced concrete or dowelling in
additional reinforcement. The technique of strengthening
concrete structures by bonding steel plates to the surface of
the tension zone with adhesives and bolts was developed in
the 1960s. Since about the late-1980s the use of fibrereinforced polymers (generally known as FRPs) in this
application has been developing rapidly.
Initial applications were with FRP plates, generally
containing carbon fibres. They have many advantages over
steel plates in this application and they can be used in
situations where it would be impossible or impractical to use
steel: for instance, they can be formed in place into
complicated shapes. Fibre-reinforced polymers are lighter in
weight than steel plates of equivalent strength. This makes
installation much simpler and quicker and eliminates the
need for temporary support for the plates in most
circumstances while the adhesive gains strength. Fibrereinforced polymers can also be easily cut to length on site.
Some types of fibre are also available in the form of fabrics,
which can be bonded to the concrete surface. The chief
advantage of fabrics over plates is that they can be wrapped
round curved surfaces, for example around columns or
completely surrounding the sides and soffits of beams.
A sketch showing a wide range of strengthening applications
to a hypothetical structure is shown in Figure 1. Clearly not
all structures are suitable for strengthening using fibre
composites. A major limitation will be when the concrete
strength is low or where there are ongoing corrosion or other
durability problems. The amount of strengthening that can
be applied will often be limited by failure being induced
elsewhere in the structure.

Figure 1: Different types of structural strengthening, applied


to beams, slabs, walls and columns.

Flexural strengthening can be achieved by bonding pultruded strips or rods into slots cut in the cover region of the
concrete. This application is termed Near-Surface-Mounted
(NSM) reinforcement, and has benefits where the exposed
concrete surface is to be trafficked or otherwise exposed to
potential damage. In the UK this technique has been applied
to car park decks, and overseas to jetties and dock-side
structures that are subjected to loads from the movement of
shipping containers. The technique is also applicable where
the surface of the concrete is undulating, or if there is
excessive laitance or a thin layer of poor quality concrete
near the surface. Installation is more costly than for
externally bonded reinforcement, due to the need to cut the
slots and a slightly more complicated method for surface
preparation. Usually the technique would only be used where
externally bonded reinforcement is not a good technical
solution.
An appreciable number of structures in the UK and
elsewhere have been strengthened using FRP materials and
the rate at which the technique is being used is increasing
rapidly. It is estimated that several hundred structures have
been strengthened in the UK to date, with the amounts of
fibre composite material involved ranging from a few metres
(or m2) for a small job to several kilometres on a major one.
There was little independent guidance on how the design of
strengthening works should be carried out until the First
Edition of Technical Report 55 was published in December
2000(1). Subsequently, guidance documents have been

Design guidance for strengthening concrete structures using fibre composite materials

published in various countries, including the USA(2) and


Canada(3). In addition, the Canadian Standards Association
has published the first national code for strengthening with
FRP(4). Following the publication of TR55, the Concrete
Society published Technical Report 57, Strengthening
concrete structures using fibre composite materials: acceptance, inspection and monitoring(5) in 2003, which covers
the vitally important topics of inspection and maintenance.
Materials and techniques are developing rapidly, as is the
range of applications. Hence it was thought necessary to
produce this Second Edition of TR55. A number of the
changes are matters of detail, brought about by additional
research findings and further experience of the use of the
materials. However, significant changes or additions have
been made in some areas, as follows:

Material and system selection guidance


The treatment of partial safety factors in the design process
has been modified; factors are now applied to the FRP
strains rather than the stresses
Extreme loadings
A more systematic and comprehensive approach to separation failure
Design of members in flexure, to give a more rational
approach that is closer to that used for reinforced concrete
Design of members in shear, to provide a less empirical
approach, allowing wider and more confident application
of the technique
Column design, to provide a more unified approach for
axial and combined axial and flexural strengthening. A
more detailed approach has been developed for the
strengthening of rectangular columns
Flow charts for flexure, separation, shear and column
design

Design guidance for the new technique of NSM reinforcement


Overview of emerging technologies, such as the use of
prestressed composites, mechanical anchorage systems
and alternatives to the adhesives currently used.

The guidance in this Report is not specific to any particular


type of FRP material or any particular strengthening
technique. It covers the use both of manufactured composite
materials bonded on or near the concrete surface and composites formed in situ on the surface.
The Report deals mainly with the design of strengthened
members, i.e. beams, slabs and columns. Other aspects, such
as currently available materials, appropriate application
techniques and current uses, are described. It is intended to
cover the principles involved, not the detailed approaches
that are applicable to individual materials and techniques.
Further details of material properties and techniques can be
obtained from materials suppliers and from specialist
designers and contractors. The important topics of inspection
and maintenance are covered briefly in this Report; more
detailed coverage is given in Concrete Society Technical
Report 57(5), which should be read in parallel with this
Report; reference is made to specific parts of Technical
Report 57 (abbreviated to TR57) where appropriate.
The Report is specifically concerned with strengthening
concrete structures. Fibre composites have been successfully
used to strengthen metallic(6) and other structures. The basic
principles of this Report will still be applicable but the
detailed design recommendations will not apply.
To help readers unfamiliar with composites, a glossary of
terms is given in Appendix A.

BACKGROUND

2.1 PRINCIPLES OF STRENGTHENING


A concrete structure may need strengthening for many
reasons. Examples are:

To increase live-load capacity, e.g. of a bridge subject to


increased vehicle loads or a building, the use of which is
to change from residential to commercial
To add reinforcement to a member that has been underdesigned or wrongly constructed
To improve seismic resistance, either by providing more
confinement to increase the strain capacity of the concrete,
or by improving continuity between members
To replace or supplement reinforcement, e.g. damaged by
impact or lost due to corrosion. (This will only be practical if the cause of the damage is identified and treated.)
To improve continuity, e.g. across joints between precast
members
To provide replacement reinforcement following structural
alterations, e.g. around holes cut through floor slabs for
lift or stair installation or through walls to accommodate
new services.

In most cases it is only practical to increase the live-load


capacity of a structure. However, in some situations it may be
possible to relieve dead load, by jacking and propping, prior
to the application of the additional reinforcement. In these
cases, the additional reinforcement will play its part in carrying the structures dead load. Prestressing techniques using
composite materials are being developed that will also help
to carry part of the dead load. This approach is discussed
although not covered in detail in the design sections of this
Report.
Three basic principles underlie the strengthening of concrete
structures using fibre composite materials, which are the same
irrespective of the type of structure:

Increase the bending moment capacity of beams and slabs


by adding fibre composite materials to the tensile face
Increase the shear capacity of beams by adding fibre
composite materials to the sides in the shear tensile zone
Increase the axial and shear capacity of columns by
wrapping fibre composite materials around the perimeter.

These forms of strengthening can be used to increase the


seismic performance of structures and also to enhance the
performance of joints between members.

2.2 ASSESSMENT OF STRUCTURES TO


BE STRENGTHENED
The decision to strengthen a structure will come at the end of
what may be a prolonged assessment process. This is illustrated, in outline only, in Figure 2. The process is independent of structure type and should be based on rigorous
criteria and sound engineering judgement. The assessment
process will usually involve some investigation of the condition of the structure or some re-analysis and study of the background issues. Guidance may be obtained from documents
such as Concrete Society Technical Report 54 Diagnosis of
deterioration in concrete structures(7) and the Institution of
Structural Engineers Appraisal of existing structures(8). In
all cases an experienced engineer should be part of the assessment team. The process will usually be aimed at providing
answers to some or all of the following questions:

Has the condition or load-carrying capacity of the structure decreased significantly?


Has the loading changed significantly?
Is the concrete of adequate quality and strength to make
strengthening a feasible option? This applies to the structure overall and not just to the surface or surfaces to
which the FRP is to be bonded. It is suggested that the
minimum concrete strength should be 20N/mm2 with a
pull-off strength of the surface in the zone to be
strengthened 1.5N/mm2.
What are the risks to the public, to commerce and to the
structure of taking no action?
What are the cost implications of strengthening, including
direct costs, future costs and the cost of disruption while
the work is carried out?
What are the cost implications of demolition and rebuilding, including direct costs, future costs, costs associated
with loss of use of the structure and disruption while the
work is carried out?
What is the anticipated future life of the structure in its
present form?
Will inspection and maintenance be possible?
How would strengthening works affect local infrastructure, commerce, safety and the environment?
Are any political issues involved?
What is the age of the structure and is it of historical
importance?
What parties and authorities would be required to approve
the works?
Are there any programming or funding constraints?

Design guidance for strengthening concrete structures using fibre composite materials

Figure 2: Flow chart of assessment process.

By addressing these issues, decisions about the appropriate


action for a particular structure can be made. In some cases
strengthening will not be a sensible option, unless remedial
work is carried out first. Examples are structures with significant materials problems, such as high chloride content
leading to severe reinforcement corrosion. In general it will
not be appropriate to strengthen a deteriorated structure
unless the cause of the deterioration (e.g. chloride ingress)
has been addressed and, where possible, mitigated.
Once it has been decided that strengthening is a realistic
option and that the structure is suitable for strengthening, the
next step is to identify an appropriate strengthening scheme.
The feasibility study should include consideration of the
points listed above in relation to possible schemes, such
issues as whole-life costs of the various options and careful
assessment of the residual life and strength of the structure.
The risks associated with each option should be assessed
during the feasibility study. This assessment should compare
the possible higher risks associated with new techniques
with little history of long-term performance to those of older,
tried and tested methods. However, the benefits of newer
techniques can outweigh this perceived disadvantage: the
risks associated with premature failure are low if
strengthening is to be provided only for the live-load case.

2.3 STRENGTHENING SOLUTIONS


Strengthening solutions considered in a feasibility study can
range from repair of a damaged structure in order to restore
its original strength to adding elements to increase its capacity.
All solutions are, to a greater or lesser extent, project-specific
but some general approaches are commonly used. Repair typically involves crack injection and/or breaking out damaged
areas and reinstating with cementitious repair mortars or flowing concrete. As stated above, this approach is used where
the aim is to restore the original strength of a structure.
The most common traditional techniques for strengthening
are as follows:

Increase the reinforced concrete cross-section. Approval


authorities and owners of structures usually readily
accept this solution as it has a proven track record.
However, loading restrictions are required while the concrete cures to an acceptable strength. This restriction may
be critical in some instances for example where a bridge
closure would lead to unacceptable disruption.
Add prestressing to relieve dead load. Like increasing the
cross-section, this technique has a proven track record

Background

and gains ready acceptance. Loading restrictions may be


required during installation, which may not be acceptable. This technique requires the existing structure to be
capable of withstanding high local prestressing forces.
Use plate bonding to enhance tensile reinforcement of
elements. Steel plate bonding has been widely used and
can be considered to have a proven track record. Design
guidance is given in the Highways Agency Advice Note
BA 30/94(9). Disadvantages of the technique are the weight
and difficulty of handling the plates, difficulty in cutting
to shape, the need to apply and maintain corrosion protection and to anchor the plates to the concrete section while
avoiding damage to embedded reinforcement. As discussed above, access and installation times may be critical
issues in some locations.
Add material to provide confinement of the concrete in
compression members. This can be achieved by installing
in situ reinforced concrete or prefabricated steel collars
or wrapping the element with resin-bonded fibre composite material. The use of collars is the most common
technique where space permits. The technique tends to be
readily accepted as the increase in the cross-section can
be clearly seen. With in situ reinforced concrete collars,
loading restrictions on the structure are required while
the concrete gains strength.
Shear strengthening. This can be achieved by installing
external steel straps to beams.

Fibre composite strengthening is seen as a viable alternative


to some of these traditional methods. Fibre composite plate
bonding is being used widely in place of steel plate bonding
because of the speed and ease of installation and the ease
with which the material can be cut to shape and bent to fit
slightly curved surfaces. As the technique is relatively new,
a proven long-term track record does not exist and this is
seen by some as a disadvantage. However, the basic technique and the adhesives are similar to those used for steel
plates, which have been widely and successfully applied.
Fibre composites are particularly attractive in locations where
space does not allow a significant increase in cross-section
or where the installation time is critical. Columns can be
strengthened by wrapping with fibre composite material, to
increase their axial capacity and their resistance to bending
and shear. Overall the advantages of fibre composites tend to
outweigh the perceived disadvantage of a lack of track record
and the reluctance of some approval authorities and owners
of structures to adopt new materials.

The lower weight makes handling and installation significantly easier than steel. This is particularly important when
installing material in cramped locations. Figure 3 shows
carbon fibre plates being installed in a culvert with limited
headroom.

Figure 3: Installing fibre composite plates in a culvert.

Work on soffits of bridges and building floor slabs can often


be carried out from man-access platforms rather than full
scaffolding. Steel plate requires heavy lifting gear and must
be held in place while the adhesive gains strength. Bolts
must be fitted through the steel plate into the parent concrete
to support the plate while the adhesive cures and to reduce
the effects of peeling at the ends. When applying FRP plate
or sheet material pressure is applied to the surface using a
roller to remove entrapped air and excess adhesive. It may be
left unsupported. In general, no bolts are required; in fact,
the majority of FRP strengthening material is uniaxial (that
is all the fibres are aligned in one direction) and the use of
bolts would seriously weaken the material unless additional
cover plates are bonded on or the plates are designed with a
proportion of fibres in the transverse direction. Furthermore,
because there is no need to drill into the structure to fix bolts
or other mechanical anchors there is no risk of damaging the
existing reinforcement. Fibre composite materials are available in very long lengths while steel plate is generally limited
to 6m. The availability of long lengths and the flexibility of
the material (see Figure 4) also simplify installation:

2.4 ADVANTAGES AND DISADVANTAGES


OF FIBRE COMPOSITE
STRENGTHENING

2.4.1 Advantages

Fibre composite strengthening materials have higher ultimate


strength and lower density than steel. When taken together
these two properties lead to fibre composites having a
strengthweight ratio significantly higher than steel plate in
some cases, though it is generally not possible to use this fully.

Laps and joints are not required.


Within limits, the material can take up irregularities in
the shape of the concrete surface.
The material can follow a curved profile; steel plate
would have to be prebent to the required radius.
The material can be readily installed behind existing
services (see Figure 5).
Overlapping, required when strengthening in two
directions, is not a problem because the material is thin
(see Figure 6), but care must be taken with the application process in the region of the overlaps.

Design guidance for strengthening concrete structures using fibre composite materials

In terms of environmental impact and sustainability, studies


have shown that the energy required to produce FRP materials
is less than that for conventional materials. Because of their
light weight, the transport of FRP materials has minimal
environmental impact.

Figure 4: Installing FRP plate, showing the flexibility of the


material.

These various factors in combination lead to a significantly


simpler and quicker strengthening process than when using
steel plate. This is particularly important for bridges because
of the high costs of lane closures and possession times on
major highways and railway lines. It has been estimated that
about 90% of the market for plate strengthening in Switzerland has been taken by carbon plate systems as a result of
these factors.
2.4.2 Disadvantages

Figure 5: FRP plate installed behind existing services.

The main disadvantage of externally strengthening structures


with fibre composite materials is the risk of fire, vandalism
or accidental damage, unless the strengthening is protected.
A particular concern for bridges over roads is the risk of soffit
reinforcement being hit by over-height vehicles (bridge
bashing). However, strengthening using plates is generally
provided to carry additional live load and the ability of the
unstrengthened structure to carry its own self-weight is unimpaired (see also Section 2.7). Damage to the plate strengthening material only reduces the overall factor of safety and
is unlikely to lead to collapse. An additional cause of damage
is that from following trades, such as drilling through FRP to
fix brackets, etc.
As detailed later, workmanship is critical to the success of a
fibre composite strengthening scheme. Thus a further cause
for concern is the difficulty of ensuring that the work is
carried out correctly. In the UK a certification scheme for
operatives and supervisors involved with strengthening has
been proposed and is currently being developed by TWI.
Problems with the adhesive layer will not generally be
visible from the surface. Similarly, it is difficult to assess the
presence of voids in wet lay-up systems and thus there is
uncertainty as to the properties that have been achieved.

Figure 6: Overlapped carbon FRP plates on Dudley Port


Bridge, West Midlands.

The materials fibres and resins are durable if correctly


specified, and require little maintenance. If they are damaged
in service, it is relatively simple to repair them, by adding an
additional layer.
The use of fibre composites does not significantly increase
the weight of the structure or the dimensions of the member.
The latter may be particularly important for bridges and other
structures with limited headroom, and for tunnels.

Currently the properties of materials used in FRP strengthening schemes are not covered by British or International
Standards. However, the Classification and Assessment of
Composite Materials Systems for use in the Civil Infrastructure project, being led by Oxford Brookes University, is
developing a classification scheme for adhesives and laminating resins. Further information may be obtained from the
web site (www.compclass.org). In addition the British Board
of Agrment is developing approval for certain types of
strengthening materials used in bridges under the Highways
Authorities Product Approval Scheme (HAPAS).
Experience of the long-term durability of fibre composites is
limited, though some installations have been in service for
13 years. This may be a disadvantage for structures for which

Background

a very long design life is required (see Section 2.5) but can
be overcome by appropriate monitoring (see Chapter 11) and
as detailed in TR57(5).

2.6 ECONOMICS

A perceived disadvantage of using FRP for strengthening is


the relatively high cost of the materials. However, comparisons should be made on the basis of the complete strengthening exercise (see Section 2.6) taking into account hidden
costs such as delays and disruptions to the users of the
structure. Installation can require large areas of the concrete
surface to be prepared, particularly with fabrics, which can
be labour-intensive.

The relative economics of the use of fibre composites and


other strengthening systems depend on the circumstances.
Many factors are involved, and it is necessary to compare
costs both in the short and long term. The latter may be
difficult to quantify as the life-time behaviour can only be
estimated fairly crudely. In many cases the alternative may
be demolition and replacement of the structure, with the
consequent disruption.

A disadvantage in the eyes of many clients will be the lack


of experience of the techniques and suitably qualified staff to
carry out the work, but this can be overcome by using
suitably qualified designers and contractors.

Factors such as the cost of access and possession time should


be taken into account as they can have a significant influence.
High closure costs are often incurred by highway and railway works. These will vary significantly depending on a
range of factors, including the location, the season and the
time of day. However, they will not take into account the
social costs of disruption. As an example, upgrading of a
major highway in New York City had to be carried out at
night as there was a requirement for the road to be fully open
during the day. The penalty for failure to reopen the carriageway in the morning was $30,000 per hour, with a penalty of
$20,000 per day for overrun of the complete project(12).

2.5 DESIGN LIFE


The Highways Agency document BD 84/02, Strengthening
of concrete bridge supports for vehicle impact using fibre
reinforced polymers(10), uses 30 years for the design life of a
fibre composite strengthening system. This figure is considered reasonable, based on current experience of the adhesives used in steel plate bonding. There is considerable
experience of the use of adhesives in other applications, such
as marine structures, which would suggest a design life of at
least 40 years. Fibre composite structures such as the West
Mill Bridge(11) have been designed for significantly longer
lives.
Ideally, the design life for the strengthening system should
be related to the remaining life of the structure and should
take into account the future plans for the structure. In many
cases, if a mature structure is to be strengthened, a 30-year
life for a strengthening system may well be appropriate.
However, this may not be the case for structures with long
design lives, such as bridges and nuclear structures. Here, it
may be necessary to accept a strengthening system with a
design life less than the anticipated remaining life of the structure, on the understanding that the life of the strengthening
system will be reassessed at a future date.
Because of the relative lack of long-term experience of the
performance of fibre composite strengthening systems,
regular inspection and maintenance regimes should be
instigated (see Chapter 5 of TR57). This is particularly
important for buildings, which, unlike bridges, are not
generally subjected to any form of routine inspection. Where
practical, additional material should be installed, which can
be removed at a later stage for testing. This approach has
been adopted on a number of structures including the Barnes
Bridge in Manchester and the John Hart Bridge in British
Columbia (see Section 4.3). It may be possible to incorporate some form of monitoring system in the fibre composite.

2.6.1 Installation

Studies carried out for Railtrack (now Network Rail) have


indicated that strengthening with FRP materials will be
approximately 30% cheaper than the equivalent strengthening
using steel plate. The use of FRP for column strengthening
on one UK highway bridge halved the cost, as well as shortening the contract duration and significantly reducing the
need for lane closures.
Loss of revenue can be significant when a structure is understrength and hence cannot be used to its full capacity. It was
reported that the Trenchard Street car park in Bristol was
losing 1M per year in lost sales prior to strengthening.
In Florida, the beamcolumn connections in a parking garage
were strengthened by bonding carbon fibre sheet material to
the sides of the beams(13). It was estimated that the adhesively
bonded repair was 35% cheaper than the conventional
method, which would have involved dowelling in additional
steel reinforcement and encasing the joint with additional
concrete.
In Edmonton, Canada, carbon fibre reinforced polymer composite sheet material was applied to the soffits and sides of a
bridge, to improve its shear resistance(14). The cost was
reported as $70,500 for strengthening the complete bridge. A
conventional external stirrup system was estimated to cost
some $100,000. Thus the bonded solution showed approximately 30% saving in costs, due chiefly to the fact that the
work was carried out from below the bridge and avoided the
traffic closures that would have been required for the
conventional system.

Design guidance for strengthening concrete structures using fibre composite materials

Beams of the Maryland Street Bridge in Winnipeg, Canada,


were strengthened with vertical and horizontal sheets of
carbon fibre to increase the shear capacity. It was estimated
that the cost was about 70% of the conventional approach,
which would have involved removing parts of the bridge
deck, installing post-tensioned external shear stirrups and
casting additional concrete round the beams. This comparison was on the basis of direct costs and did not consider
factors such as traffic delays.
A 30-year-old processing tower in Qatar was strengthened
with 3500m of carbon FRP plate. Several options were considered but the material was chosen because of the speed of
installation. The plant was shut down for 25 days to allow the
work to be undertaken, but the contract was actually completed in 20 days. This enabled production to restart earlier than
planned, which was clearly of great benefit to the operators
of the plant(15).
Hooks and Cooper(16) give two examples of significant cost
savings. The crossheads of a 1950s bridge in New York were
strengthened in flexure and shear using carbon FRP plates, at
a cost of $18,000. It was estimated that conventional repair
would have cost $150,000. Concrete box beams in Kentucky
were strengthened at a cost of $105,000, when replacement
of the structure would have cost $450,000.
Some economic considerations for particular applications are
reported in later chapters. Unfortunately, the information is
largely qualitative, but can be used for guidance when investigating the economics of a situation.
2.6.2 Whole-life costing
The technique of whole-life costing can play an important part
in making decisions on when and how to repair or strengthen
concrete structures. This is recognised in BS EN 1504 Part
9(17), which lists among the factors to be considered when
choosing between repair options:

The number and cost of repair cycles acceptable during


the design life of the concrete structure
The cost and funding of the alternative protection or repair
options, including future maintenance and access costs.

The whole-life cost of a repair or strengthening solution is


the sum of the initial (installation) cost and the future
(maintenance) costs over the remaining life of the structure.
To permit meaningful comparisons to be made, future costs
are discounted to present day value. To carry out a life-cycle
cost analysis requires an understanding of:

deterioration processes as they relate to the particular


structure or different parts of the structure
repair and strengthening methods and their durability
costs of repair or strengthening and maintenance activities
indirect costs due to loss of service
the owners requirements for the serviceability and service
life of the structure.

In many cases, the basic data to permit reasonable assessments of the various elements that make up the whole-life
cost are not available. Nonetheless, it can be appreciated that
strengthening using fibre composites can be competitive in
whole-life cost comparisons because both installation and
maintenance costs are usually lower than those of competing
techniques and possession times are shorter.
Prolonging the useful life of structures that will still be
required for a long time into the future (e.g. road or rail
bridges) becomes an attractive proposition in whole-life cost
terms. This is because, if replacement can be delayed for
many years, the cost at present day value is considerably
reduced. For example, if a discount rate of 8% is assumed, a
cost of 1,000,000 at year 20 has a present day value of only
200,000. It can be more economic, in whole-life cost terms,
to strengthen now and replace in 20 years, than to replace now.
One factor which is difficult to take into account in wholelife costing is the time until the structure becomes obsolete.
This may happen for physical, economic, functional, technological, social or legal reasons. This uncertainty can lead to
the lowest initial cost option being favoured on the basis that
there is little to be gained from additional spending now, if
the structure is unlikely to be required in its present form in
ten years.

2.7 LEVEL OF STRENGTHENING


A key factor in the choice of strengthening system will be the
level of strengthening (i.e. the maximum increase in load
capacity) that can be achieved. Strengthening against one
mode of failure (e.g. bending) may increase the probability
of occurrence of another mode (e.g. shear). This must be
considered in the design process. In addition, the design must
explicitly consider the risks associated with any possible
partial or complete failure of the strengthening, due for
example to fire, vandalism or accidental damage. Because of
the lack of long-term experience of fibre composite strengthening, some clients are recommending that the approach
should only be used to increase the factor of safety against
collapse. In other words, failure of the composite will not
lead to the collapse of the structure.

MATERIAL TYPES AND PROPERTIES

3.1 INTRODUCTION
Fibre composites are formed from high performance fibres
combined with an appropriate resin. Epoxies are generally
used, but some development has been carried out on inorganic
cement-based matrices(18). For strengthening applications,
the composite may be preformed into plates or panels and
bonded to the concrete. The most common example is
composite plates bonded to the soffits of beams or slabs.
Alternatively, the fibres may be combined with the resin in
situ as part of the application process, such as in the wrapping of columns. The mechanical properties of fibre composites are chiefly controlled by the type, amount, orientation
and distribution of fibres in the cross-section. The role of the
resin is to transfer stresses to and from the fibres and also to
provide some protection from the environment. This Chapter
provides a general introduction to the fibres and resins used
for strengthening. For further information on the properties
and behaviour of composites, the reader should consult
standard textbooks, such as An introduction to composite
materials(19) and Composite materials: engineering and
science(20).

3.2 FIBRES
3.2.1 Types of fibre
The most suitable fibres for strengthening applications are
glass, carbon or aramid. (Aramids are better known by the
trade names Kevlar and Twaron.) Each is a family of fibre
types in general, with individual fibre types within the families that may vary. Typical values for the properties of fibres
are given in Table 1. It should be noted that these values are

for the plain fibres alone, not woven fabrics nor for the resulting fibre composites. The strength and modulus for manufactured composites will be significantly lower (see Sections
3.4 and 3.6). The values in Table 1 should only be taken as
indicative; where necessary, actual values should be obtained
from the manufacturer. The fibres all have a linear elastic
response up to ultimate load, with no significant yielding.
Details of some available materials are given in Table B1 of
Appendix B.
3.2.2 Performance of different types of fibre
The selection of the type of fibre to use in a particular
application will depend on many factors the type of structure, the expected loading, the environmental conditions, and
so on. Some information is given in this section; further
advice can be obtained from the suppliers of strengthening
materials. Throughout, the comments refer to the performance
of the fibre itself; in most situations this will be modified by
the resin or adhesive.
Chemical resistance
Carbon and aramid fibres are resistant to most forms of
chemical attack. Many types of glass fibre, including the
widely used E glass, are attacked by alkalies (pH greater than
about 11) but not by acids. Alkali-resistant (AR) glass fibres
are specially formulated for use in highly alkaline environments and are therefore suitable for strengthening concrete
structures. Aramids absorb much more water than either of
the other two fibres, which can cause problems with the
resinfibre interface. There is some evidence to suggest that,
in the presence of salts, fracture of all types of fibre can occur
due to the formation of angular crystals.

Table 1: Typical dry fibre properties.

Fibre
Carbon: high strength*
Carbon: high modulus*
Carbon: ultra high modulus

Tensile strength
(N/mm2)
43004900
27405490
26004020

Modulus of elasticity
(kN/mm2)
230240
294329
540640

Elongation (%)

Specific density

1.92.1
0.71.9
0.40.8

1.8
1.781.81
1.912.12

Aramid: high strength and high modulus 32003600

124130

2.4

1.44

Glass

7085

3.54.7

2.6

24003500

* Based on polyacryonitrile precursor


Based on pitch precursor
Aramids with the same strength but a lower modulus are available but are not used in structural strengthening applications.

Design guidance for strengthening concrete structures using fibre composite materials

Resistance to ultraviolet light

Stiffness

Glass and carbon fibres are not affected by ultraviolet light.


Aramid fibres change colour under ultraviolet light and the
strength is reduced. However, when embedded in a resin
matrix this degradation only occurs near the outer surface
and there is little effect on the overall mechanical properties.
(Direct exposure to sunlight can embrittle all resins and a
protective paint is normally recommended if direct exposure
is likely.)

The elastic modulus of carbon fibre is similar to, or significantly greater than, that of steel. The stiffness of aramid is
lower and that of glass significantly lower.

Electrical conductivity
Aramid and glass fibres are non-conducting and hence are
suitable for use close to power lines, electrified railway lines
and communications facilities. As carbon fibres conduct electricity they should be electrically isolated from any steel to
prevent the establishment of a galvanic cell. In general the
resin will be sufficient for this, but where there is a particular
risk it is recommended that a glass fibre sheet be additionally
included as the outermost layer of the FRP strengthening
system.
Designers should also be alert to the possibility of carbon
fibres within an FRP attracting induced currents when placed
close to an AC electricity supply. While no experimental
work appears to have been carried out in this area, it is theoretically possible that induced currents within a carbon FRP
could lead to unacceptable heating of an ambient cure adhesive as it has been shown that the conducting properties of
carbon fibre can be used to pass an electric current to achieve
a higher adhesive cure temperature.
For UK railway applications it is a requirement that any
conducting material that could become live due to induced
currents or short circuits from traction power sources must be
electrically connected to the return conductor. For metallic
structures this is normally achieved by attaching an electrical
bond between the return conductor and the structure. Due to
the distributed nature of carbon fibres within the adhesive
matrix of a carbon FRP it is virtually impossible to guarantee
that every single fibre can be connected to the return conductor by an electrical bond. Hence Network Rail will only
permit the use of aramid FRP in close proximity to its AC
overhead electrification systems; however, carbon FRP is
permitted where DC electrification systems are present.
Care is needed when handling or cutting carbon FRP close to
electrical equipment due to the risk of short-circuit by airborne particles (see Section 3.9). In addition, when used close
to power lines etc., steps must be taken to ensure that, in the
unlikely event of adhesive failure, the composite does not
come into contact with the electrical source.
Compressive strength
The compressive strengths of carbon and glass fibres are close
to their tensile strengths; that of aramid is significantly lower.

10

Impact resistance
Performance of fibres during impact is highly dependent on
the elastic strain energy generated and absorbed. Fibres combining high strength with high elongation (tensile strength
greater than 3,500N/mm2 and elongation greater than 2%)
are most suitable for applications where impact resistance is
important. Selected grades of carbon, aramid and glass fibre
can meet these requirements.
Fire
Glass fibres retain strength up to their melting point (over
1000C) while carbon fibres oxidise in air above 650C.
Aramid fibres are not normally used above 200C. None of
the fibres will support combustion. In composites, the resin
behaviour will dominate performance; most generate toxic
smoke. Several composite systems have coatings that can
provide protection.

3.3 FABRICS
Fabrics are available in two basic forms:

Sheet material. The fibres are generally in a unidirectional arrangement, though biaxial and triaxial arrangements are available. They may be on a removable backing
sheet or in the form of a woven or stitched cloth.
Prepreg material. This consists of fibres preimpregnated
with resin, which is cured once in place, by the application
of heat or by other means.

The selection of the appropriate form of fabric will depend on


the application.
The properties of the sheet materials depend on the amount
and type of fibre used. An additional consideration is the
arrangement of the fibres; parallel lay gives unidirectional
properties while a woven fabric has bidirectional properties.
In woven fabrics, perhaps 70% of the fibres are in the strong
direction and 30% in the transverse direction. It should be
noted that the kinking of the fibres in the woven material
significantly reduces the strength and stiffness. In addition
some fabrics are formed with equal amounts of fibres in two
directions, at 45 to the longitudinal axis.
The thickness of the material will depend on the type and
arrangement of the fibre. Fabrics are available in various
widths to suit the particular application. The sizes and properties of some available materials are given in Table B2 of
Appendix B.

Material types and properties

3.4 PLATES
Unidirectional plates are usually formed by the pultrusion
process. Fibres, in the form of continuous rovings, are drawn
off in a carefully controlled pattern through a resin bath,
which impregnates the fibre bundle. They are then pulled
through a die, which consolidates the fibreresin combination and forms the required shape. The die is heated which
sets and cures the resin, allowing the completed composite to
be drawn off by reciprocating clamps or a tension device.
The process enables a high proportion of fibres (generally
about 65%) to be incorporated in the cross-section. Hence, in
the longitudinal direction, relatively high strength and stiffness are achieved, approximately 65% of the relevant figures
in Table 1. Because most of, if not all, the fibres are in the
longitudinal direction, the transverse strength will be very low.
Plates formed by pultrusion are 14mm thick and are
supplied in a variety of widths, typically between 50 and
150mm. The sizes and properties of some available plates
are given in Table B3 of Appendix B. (It should be noted
that, while plate properties and dimensions of plates can be
tailored to suit the particular application, it will generally be
more economical to use stock sizes.) Carbon is the most
widely used fibre though glass is used in some applications.
As pultrusion is a continuous process, very long lengths of
material are available. Thinner material is provided in the
form of a coil, with a diameter of about 1m, as shown in
Figure 7. It can be easily cut to length on site using a simple
guillotine (see Figure 8).

can be produced, with the width and thickness being tailored


to the particular application. Widths up to 1.25m have been
produced and thicknesses up to 30mm. Other forms of manufacture, such as resin infusion, are sometimes used but these
are generally less attractive commercially.

Figure 8: Cutting carbon FRP plate on site.

3.5 RODS AND STRIPS FOR NEARSURFACE-MOUNTED (NSM)


REINFORCEMENT
As NSM material is installed within the cover region of the
concrete, the diameter of the rod, or maximum possible
dimension of the cross-section of the strip, is limited. Most
experimentation to date has used circular bars of diameters
in the range 716mm, or rectangular strips of thickness less
than 2mm. Initial UK applications have used carbon FRP bar
with a circular cross-section of less than 10mm diameter.
The sizes and properties of some rods and strips are given in
Table B4 in Appendix B.

3.6 PREFORMED SHELLS FOR COLUMN


CONFINEMENT

Figure 7: Coil of carbon FRP plate.

Plates can also be produced using the prepreg process, which


is widely used to produce components for the aerospace and
automotive industries. Typically plates have a fibre volume
fraction of 55% and can incorporate 10% off-axis fibres
(usually glass aligned at an angle of 45 to the longitudinal
axis) to improve the handling strength. Lengths up to 12m

Preformed shells have been used to strengthen columns on a


number of structures. (It should be noted that the basic
principles of strengthening columns given in Chapter 8 are
applicable, but strengthening with shells is a more complex
design process.) There have been a number of applications in
North America but only one in the UK to date. For a circular
column, the most appropriate manufacturing process is
probably filament winding. Resin-impregnated fibres are
wound round a mandrel, in the pattern required to give the
required hoop and longitudinal properties. Once fully cured,
the cylindrical shell is removed from the mandrel and cut
longitudinally so that it can be bonded round the column.
Alternatively, shells can be formed, by hand lay-up or other
processes, on the inside or outside of a suitable mould.

11

Design guidance for strengthening concrete structures using fibre composite materials

In general, the internal diameter of the shell should be close


to that of the external diameter of the column, to keep the
increase in the overall diameter to a minimum. Typically,
shells are installed with a clearance of between 50 and 150mm
from the concrete surface, with the annulus later filled with
an expansive grout. This will induce a permanent tensile stress
in the composite and compression in the concrete. It will be
necessary to check that the stress in the FRP is low enough
to avoid the risk of stress rupture.

(usually solvent-free, two-pack materials which cure at


ambient temperature). The properties of some available
epoxy adhesives and laminating resins are given in Tables
B5 and B6 in Appendix B. Generally the adhesives should be
procured from the same supplier as the plates or fabrics, to
ensure that the materials are compatible.

The strength and stiffness of the shell in the hoop and


vertical directions will depend on the type and proportion of
fibres in the cross-section and on the method of manufacture
of the composite. They will be significantly lower than the
values in Table 1. The performance of the shell is highly
dependent on the efficiency of the connection between the
component FRP units.

Because of the cost of fabricating mandrels or moulds, this


approach is only likely to be cost-effective when a large
number of identical columns are being strengthened, such as
in multi-span bridges or multi-storey buildings.

The selection of the type of epoxy to be used in a particular


application is governed by various factors, including the
environment and the required speed of fabrication. Advice
should be obtained from the adhesive manufacturer. The adhesive should be able to withstand a maximum temperature of
50C in service and generally have a glass transition temperature (Tg) between 50 and 65C. In special circumstances,
such as bonding FRP material to the top surface of a bridge
deck which is to receive hot bituminous surfacing (see Section
4.3.1), the adhesive may be heated significantly. This may
require an epoxy to be selected with a higher glass transition
temperature but the adhesives performance at lower temperatures may then be affected. Advice should be sought from
the supplier.

3.7 SPECIALS
Plates formed into an L shape may be used as an external
link to provide shear reinforcement on beams, with the lower
leg of the L providing the anchorage for the vertical portion(21,22) (see Figure 9). The same type of unit could be used
to provide anchorage at the top of a beam, at the interface
with the slab or at beamcolumn connections. There have
been various applications of this type in Germany and
Denmark but only one in the UK.

The adhesives that are sometimes considered as alternatives


to epoxies have certain drawbacks:

Polyester adhesives have high curing shrinkage, high coefficient of thermal expansion, can be subject to alkaline
hydrolysis, and are difficult to bond to when hardened.
Vinyl ester adhesives are subject to curing shrinkage, and
the bond is badly affected by moisture.
Polyurethane adhesives have high curing shrinkage, can
be affected by moisture and are difficult to bond to.

Where fire is a significant design consideration, such as in


tunnels and confined spaces, the adhesive selected should be
one that releases a minimum amount of toxic gases. Owners
may have their own standards for the approval of materials
(e.g. Fire safety performance of materials used in the Underground(25)). Advice should be sought from the supplier.
Adhesives are generally specified on the basis that the
concrete surface is maintained in a dry condition during the
strengthening work and is in a normal atmospheric exposure
situation in service. Where the concrete surface cannot be
kept dry during the work or where, for example, the surface
is submerged or sometimes submerged in service, adhesives
with special properties may be required and specialist advice
should be obtained from adhesive manufacturers.
Figure 9: Shear reinforcement straps.

3.8 ADHESIVES AND LAMINATING


RESINS
General information on adhesives may be found in publications such as Adhesives in civil engineering(23) and A guide to
the structural use of adhesives(24). The adhesives and laminating resins most commonly used with concrete are epoxies

12

Where it is exposed to significant ultraviolet light, protective


paints, which must be compatible with the adhesive, will
generally be required to prevent the exposed epoxy resin in
a fabric system degrading. Guidance should be sought from
the supplier of the strengthening system.
For porous surfaces, a priming coat may be required, which
must be compatible with the adhesive. (It was noted earlier
that bonding to a honeycombed surface is not feasible.) As

Material types and properties

indicated in Section 10.2, the quality of the surface should be


assessed after priming by pull-off tests; tests have shown that
correctly specified primers increase the pull-off strength by
about 10%.

3.9 ENVIRONMENTAL ASPECTS AND


HEALTH AND SAFETY
3.9.1 Environmental aspects
Under CDM Regulations(26), designers in the UK must
consider all environmental aspects, including the eventual
disposal of the materials used. Aramid, glass and carbon fibres
are all non-toxic and inert, and are not considered to be
hazardous as waste. For landfill disposal, they do not contain
any substance that could leach out to contaminate the groundwater or the air. The most commonly used adhesive and
matrix materials, when fully cured, are also substantially
inert at normal ambient temperatures and so are not hazardous. However, incineration of matrix and adhesive materials
may not be an appropriate disposal method unless special
care is taken. In addition, incineration of carbon materials
may release fine electrically-conductive particles into the air.
Various approaches are being developed for recycling composites, mainly involving grinding the material to form a filler
in new composites.
3.9.2 Health and Safety
All fibres when encapsulated in cured matrix or adhesive
present negligible risk to human health in normal use. However, care must be taken when cutting and machining all
composites, because fine fibre particles may irritate skin,
eyes and mucous membranes. In addition, care must be
taken when handling resins; suitable protective clothing
should be worn. Reference should be made to the COSHH
Regulations(27) and to manufacturers data sheets. See also
Section 10.1.

3.10 CHOICE OF MATERIALS FOR


DESIGN
3.10.1 Plates versus wet lay-up sheet systems

However, there are specific instances where the use of wet


lay-up sheets is preferred over the use of plates for flexural
strengthening, often due to the lowering of longitudinal shear
stress in the adhesive layer due to the sheets being thinner
than plates. In particular, one might consider the use of wet
lay-up sheets under the following circumstances:

The working practices of the installer may dictate whether a


plate or wet lay-up system is used. From the point of view of
the environment, plates must be wiped down with a solvent
prior to installation, whereas sheets require no such chemical
preparation. On the other hand, the adhesive associated with
plates does not drip, whereas wet lay-up adhesive may drip.
While quality control needs to be particularly high during
installation of either plate or wet lay-up systems, it is fair to
say that quality control needs to be even higher for wet layup in order to minimise unevenness, misalignment, lamination defects, voids and crimping.
In situations where wet lay-up sheets are used to strengthen
structures in shear, as much of the structure should be wrapped
as possible, so that the use of sheets is preferred over that of
plates under such circumstances. Such applications might
mean U-wrapping a beam, for instance, rather than merely
adhering sheets to the sides of the beam. Furthermore, in
such circumstances, practicality of detailing means that the
U-wrap will lead to sheets aligned vertically, rather than
inclined sheets.
3.10.2 NSM systems
The following situations lend themselves to consideration of
NSM systems for strengthening:

Presently, in most concrete flexural strengthening projects,


the chosen system involves the use of carbon FRP pultruded
plates, bonded to the concrete structure through adhesive,
because:

minor unevenness in the surface can easily be bridged by


the adhesive layer of a plate system
less surface area of concrete needs to be prepared than
would be the case if wider, but thinner, wet lay-up sheets
were used
plates are usually easier to install than sheets
pultruded plates contain more fibres than a wet lay-up
sheet of similar cross-section.

high demand on longitudinal shear stress within the adhesive layer, particularly in short-span situations
poor quality substrate material, so that longitudinal shear
capacity is low
requirement for a special anchorage system, such as that
described in Section 9.1.2
strengthening around a corner
transportation of discrete plates difficult
shallow structure requiring low levels of strengthening
distributed over a large area.

The strengthened surface of the structure is trafficked or


susceptible to damage.
A thin layer of poor quality or loose concrete exists on
the surface to be strengthened, but the rest of the substrate is of high strength.
The surface is very uneven.
There is limited headroom (although installing NSM overhead can be difficult).

Particular care should be taken to prevent damage to existing


reinforcing bars when cutting the required slots. It would be
unwise to use NSM in a situation where the depth of cover
was low.

13

Design guidance for strengthening concrete structures using fibre composite materials

While NSM has been proven to be practical and of real


benefit in niche applications, its relative cost against the more
conventional plate or sheet systems should be considered,
together with the level of NSM experience in the industry,
and the availability and quality of trained NSM installers.

3.10.3 Specific composite material


To date, most concrete strengthening applications involving
composites have used a carbon system, mainly due to high
installed stiffness and strength requirements. Furthermore,
such carbon systems are usually less expensive than other
systems due to less material being required, the area of surface preparation being small and the time of installation
being short. This is also reflected in the level of worldwide
research and testing, which focuses heavily on carbon.
Therefore, it seems sensible that a carbon strengthening
system should be considered initially because of confidence
and knowledge in its use, although various reasons may
sway the choice towards other materials instead. Such circumstances where other materials (aramid or glass) should be
considered include the following:

3.10.4 Stiffness issues


When using carbon systems, it is usual to use standard
modulus fibres. Such materials are normally adequate for the
majority of strengthening schemes. Higher stiffness
materials (usually denoted HM for High Modulus) are
substantially more expensive than the equivalent standard
modulus materials, so that good reasons for their use are
usually required. Such reasons might include the following:

14

Strengthening against blast: Du Pont has conducted much


research into the use of aramid systems for such strengthening, so that the knowledge base is high.
Electromagnetically inert material is required, perhaps
near to overhead electrification on railway lines or radio/
radar installations.

Robustness and/or toughness of the material is a particularly important design criteria: under such circumstances,
aramid might be considered (although a protective layer
on carbon can be used, e.g. an abrasion-resistant layer on
a car park column).
Low-level strengthening required, so that relatively lowcost glass could be considered, placed in substantially
thicker layers than the equivalent carbon.
Wrapping of columns in the hoop direction to enhance
confinement in the event of seismic actions. Under such
circumstances, glass could be considered.

High strains cannot be induced into the carbon FRP, so


that high stiffness fibres are required.
The quantity of standard modulus carbon fibre required
for a particular stiffness is excessive.

Outside the area of concrete strengthening, it is usual to


strengthen iron and steel structures using very high-modulus
carbon FRP plates due to the otherwise large quantities of
standard-modulus carbon FRP that would be required.

REVIEW OF APPLICATIONS

4.1 INTRODUCTION
It is estimated that to date (Summer 2004), approximately
150 structures in the UK have been strengthened with FRP.
Some examples are given in Tables 2 and 3, for buildings
and bridges, respectively. The tables give details of only

those structures that have been described in published papers


or articles, i.e. for which information is in the public domain.
Details of other projects may be found in information sheets
produced by suppliers and specialist consultants. Further
details of the structures in the tables, and others, are given in
the subsequent Sections.

Table 2: Examples of strengthening of buildings in the UK.

Location
Kings College Hospital, London

Date
1996

Details of strengthening
Soffit of slab to carry
additional storey

Wormsley Library, Oxfordshire

1996

Soffit of slab, following


Carbon plate
removal of load-bearing wall

Gold and Martin(30)

Nestl Factory, Tutbury, Staffordshire

1997

Soffits of beams

Carbon plate

Luke(31), Hollaway and


Leeming(29), Taylor et
al.(32)

Allders Department Stores, Croydon and 1997


Portsmouth

Soffit of slab around


openings for new escalators

Carbon plate

Gold and Martin(30)

Nuclear power plant

Walls, to resist accidental


loading

Carbon plate

Garden(33)

Soffits of beams

Carbon plate

Luke(34)

1997

Abertillery Leisure Centre, Gwent, Wales

Material
Carbon plate

Published reference
Parker(28), Hollaway and
Leeming(29)

Car park, Manchester

2002

Columns, so that two further Carbon sheet


storeys could be added

Russell & Lomax(35),


Russell & Modi(36)

Car park, Bristol

2002

Top surface of slabs

Farmer(37)

Car park, Liverpool

2003

Top surface of slabs

Near-surfacemounted rods
Near-surfacemounted rods

Farmer(38)

Table 3: Examples of strengthening of bridges in the UK.

Location

Date

Details of strengthening

Material

Published reference

Haversham Bridge, Milton Keynes

1996

Top of slab, longitudinally

Carbon plate

Soudain(39), Anon(40),
Luke(34), Taylor et al.(32)

Subways, Jarrow, Tyne and Wear

1996

Soffit of slab

Carbon plate

Hollaway and
Leeming(29)

Underpass A413 Great Missenden, Bucks 1997

Soffit of slab

Carbon plate

Anon(41)

Devonshire Place Bridge, Skipton, North 1997


Yorkshire

Edge of slab

Carbon plate

Smith(42), Lane et al.(43),


Taylor et al.(32)

Bible Christian Bridge, A30 Bodmin


Bypass, Cornwall

Wrapping of columns

Carbon sheet
Aramid sheet
Glass sheet

Parker(44)

Greenbridge Subway, Swindon, Wiltshire 1998

Soffit of slab

Carbon sheet

Anon(45)

Glade Bridge, Leatherhead, Surrey

Soffit of beams

Carbon plate

Farmer(46)

1998

1998

15

Design guidance for strengthening concrete structures using fibre composite materials

Table 3: Examples of strengthening of bridges in the UK (continued).

Location

Date

Details of strengthening

Material

Published reference

St. Columb Gas Works Bridge, Cornwall 1998

Soffit of slab

Carbon plate

Anon(47)

Dudley Port Bridge, Dudley, West


Midlands
River Gardens Bridge, Hounslow,
Middlesex
Barnes Bridge, Manchester

1998

Soffit of slab

Carbon plate

Anon(40)

1999

Soffit of slab

Carbon plate

Barton(48)

1999

Soffit of slab

Carbon plate

Sadka(49)

Coopersale Lane Bridge, Essex

2000

Wrapping of columns

Aramid fabric

Denton et al.(50)

A19, Tyneside

2000

Wrapping of columns

Brockley slip road, M1

2001

Glass-reinforced Kendall(51), Pinzelli(52)


shell
Carbon plate
Luke and Canning(53)

Chilthurst Bridge

2002

Denton(54)

A92 Tay Road Bridge, north approach


viaduct
Patchway Viaduct, A38

2003

Upper surface of cantilever


deck
Soffit of slab with innovative Carbon fabric
anchorage
Wrapping of columns
Aramid sheet

2003

Wrapping of columns

Aramid sheet

Richardson(56)

Theydon Bois Viaduct

2002

Carbon plate

Luke and Canning(53)

St Michaels Road Bridge, Liverpool

2003

Upper surface of deck at


supports
Soffit of slab

Carbon plate

Luke and Canning(53)

4.2 BUILDINGS
4.2.1 Beams and slabs
Additional load capacity
Carbon FRP plates were bonded to the soffit of the concrete
trough slab which formed the roof of Normanby College,
part of Kings College Hospital in London, to strengthen it
sufficiently to carry an additional floor(28). It was suggested
that the conventional strengthening approach using steel
plates would not have been possible because of the problems
of inserting bolts into the soffits of the thin ribs.

Drewett(55)

to allow additional services to pass through. (Note that,


although this technique is widely used to strengthen slabs
locally, no specific guidance relating to the global performance of the slab is given in this Report. An example of the
technique is shown in Figure 10.) In addition there have been
situations where the floor slab has been strengthened with
carbon fibre sheet material rather than plates, such as the
Beyer Building, Manchester University, where openings were
formed for new air conditioning ducts.

Similarly the main beams supporting the floors in a factory


in Tutbury were strengthened using carbon fibre plates to
increase the flexural capacity by 30% to cater for the installation of new plant and processing equipment(32). The work was
carried out with minimum disruption to the factory operations.
Structural alterations
As part of the refurbishment of Allders Department Store in
Croydon, new escalators were required. This necessitated
cutting holes up to 10m by 6m in the 300mm-thick flat slab
and strengthening the adjacent slabs. After considering
various options, carbon fibre plate bonding was selected as it
minimised disruption to the operation of the store. The same
approach was used at the companys store in Portsmouth,
where new stairwells were constructed(30). Figure 10 shows
carbon fibre plates installed around a hole cut through a slab

16

Figure 10: Strengthening around hole cut through slab.

At Wormsley Library, Oxfordshire, the installation of new


services required the removal of a load-bearing wall. The
concrete slab above was strengthened with carbon fibre plate
to carry the resulting increase in dead and live loading(30).

Review of applications

Some of the main beams in a car park in Cleveland, Ohio


required modification to increase the headroom. This required
the removal of as much as 270mm of concrete from the
soffits in some areas and the installation of new prestressing
strands. Carbon fibre NSM rods were installed in the top of
the slab near the columns to help achieve the required
moment capacity(57).
Due to a design error, the simply supported beams of a warehouse in Belgium were under-strength and had insufficient
bearing at the intermediate supports. To improve the performance they were made continuous using a combination of
carbon fibre sheet and steel plate(58).
Insufficient reinforcement
Carbon FRP plates have been used to strengthen balcony
slabs in Germany to overcome problems of deflections caused
by insufficient steel reinforcement(59). The same approach has
been used in Italy.
The shear capacity of the ends of precast prestressed doubletee beams in a multi-storey car park at Pittsburgh International Airport were strengthened using carbon fibre sheet(60).
NSM reinforcement has been used to improve the seismic
resistance of a number of concrete shear walls in buildings
in Turkey. The approach has also been used for seismic
upgrading in Italy and USA.
Incorrectly located reinforcement
The top reinforcement in the cantilever slabs of a car park in
Bristol had been depressed, resulting in a cover of up to
95mm, significantly reducing the strength. Carbon fibre NSM
rods were installed to a depth of about 20mm(37) to reinstate
the strength.
At Yarborough School, Lincoln, carbon fibre strips were used
to strengthen precast stair treads which had been installed
the wrong way up.
Structural damage
In Italy carbon fibre strips have been bonded in two
directions to both faces of a prestressed double curvature
concrete shell roof structure. The structure had been
damaged, resulting in the loss of some prestress; conventional repair techniques were deemed to be not appropriate.
Carbon fibre strips were also used to strengthen the main
roof beams of an exhibition building, increasing both the
flexural and shear capacity. The ground floor beams of a
residential building, which had been damaged by an earthquake were repaired with carbon fibre sheets wrapped round
and bonded to the concrete.

Fire damage
A number of prestressed concrete beams in a multi-storey
car park in Orpington were damaged due to a vehicle fire.
Following repairs to the concrete, the beams were
strengthened with carbon FRP plates. After strengthening the
beams were load tested and insulation boards fitted to
provide one-hour fire protection. Similar work was undertaken at a retail premises in Portsmouth, using a combination
of carbon FRP plate and wrapping.
Repair
The steelconcrete composite slab of a car park in Chicago
suffered severe corrosion damage, both to the steel decking
and the top continuity steel over the supports, because of deicing salts. As part of the repair, carbon fibre plates were
bonded to the top surface of the slab over the supports. As
some of the existing cracks were about 3mm wide, the plates
were debonded on either side of the support to reduce the
peak stresses(61).
Corrosion induced by de-icing salts had seriously weakened
the decks of a car park in Liverpool. After making good the
damaged concrete, the slabs were strengthened using NSM
carbon fibre composite rods(38).
4.2.2 Columns
Wrapping a column with fibre composite (glass, carbon or
aramid) significantly increases the structural capacity of the
column. This is most effective on circular columns, and is
significantly less effective for square or rectangular columns.
Much work has been carried out in Japan and the USA with
the aim of developing cost-effective retrofitting to increase
the seismic resistance of columns. A major programme on
the performance of concrete columns enclosed by composites was carried out at Southampton University(62).
Additional load capacity
Aramid fibre sheets were used to strengthen the main
columns of a seven-level car park in Manchester so that two
further storeys could be added, providing an additional 300
car parking spaces(35,36). The material was chosen in
preference to conventional approaches, such as casting an
additional layer of concrete round the columns, because of
the speed of installation and the minimal increase in the
column dimensions.
Insufficient reinforcement
Newly constructed circular columns for a multi-storey
building in Dublin were found to have insufficient links.
They were strengthened by wrapping with carbon fibre sheet.
This approach caused minimal disruption to the construction
programme.

17

Design guidance for strengthening concrete structures using fibre composite materials

Incorrect detailing

4.3 BRIDGES

During remedial work on a multi-storey car park in West


London, it was found that the links in the columns were
located inside the main bars rather than outside. To rectify
this fault, all 400 columns were wrapped with carbon fibre
sheet in discrete bands, replicating the links.

4.3.1 Beams and slabs

Incorrect design
Excessive ground movements and floor loadings led to the
shear failure of newly-constructed square columns in the
basement car park of a hotel in Dublin. After repairing the
shear failure, the columns were strengthened by wrapping
them with carbon fibre sheet. The approach was found to be
quicker than traditional strengthening methods.

Additional load capacity


In 1997, a small concrete underpass beneath the A413 at Great
Missenden in Buckinghamshire (Figure 11) was strengthened with carbon fibre composite plates(40). The alternative
would have been the complete reconstruction of the bridge,
with consequent major traffic delays and disruption.

Additional seismic capacity


In Canada glass FRP shells have been bonded to the surface
of damaged columns to improve their load-carrying capacity.
In Japan and the USA columns have been strengthened
following earthquake damage by wrapping them with carbon
FRP, in the form of either thin strips or sheets. Similarly,
columns have been strengthened by wrapping them with
aramid fibre tape, bonded to the surface.
4.2.3 Connections
In Florida the beamcolumn connections in the parking
garage of the Palm Beach Hilton Hotel have been strengthened by bonding carbon fibre sheet material to the sides of
the beams(13). This approach was chosen in preference to the
conventional solution of increasing the size of the connection
by dowelling in additional steel reinforcement and encasing
the joint with additional concrete. It was estimated that the
adhesively bonded repair was 35% cheaper than the conventional method.

Figure 11: Underpass, Great Missenden.

The late 1960s Greenbridge Subway in Swindon was


strengthened with carbon fibre fabric to increase its flexural
capacity to that required for current traffic loadings; Figure
12 shows the material being applied. Carbon fibre plates
were used to strengthen the soffit of the River Gardens
Bridge in Hounslow(48) to increase the live-load capacity and
allow heavy vehicles into an industrial estate. Two bridges in
Crawley were strengthened with carbon fibre plates applied
to the soffits to increase the load capacity.

4.2.4 Walls
In 1997, pultruded carbon fibre plates were installed for the
first time in an operating nuclear power station in the UK(33).
The plates, of only 1m length, were bonded in several locations across structural cracks in reinforced concrete walls.
The objective was to restore the original reinforcement
contribution of the embedded reinforcing bars, which had
yielded due to widening of the cracks. The length of the
composite plates, and their cross-sectional dimensions, were
tailored to suit the substrate material properties and
anticipated design loads in the walls.
Trials in the UK and the USA have demonstrated that aramid
fibres bonded to the faces of concrete walls can significantly
increase their blast resistance.

18

Figure 12: Applying carbon fibre sheet to Greenbridge


Subway, Swindon.

Carbon fibre plates have been applied to the top surfaces of


several bridges to increase the transverse bending capacity,
including the Adur Viaduct on the A27 in Sussex and part of
the M40 in Buckinghamshire.

Review of applications

Both the bridge deck and the cross-heads of the A71


Williamston Interchange Bridge in West Lothian were found
to be under-strength. They were strengthened with two
widths of carbon FRP, which caused minimal disruption to
traffic.
The Glade Bridge (Figure 13) carries an access road over the
railway between Leatherhead and Bookham in Surrey. The
precast concrete slabs were strengthened using carbon fibre
plates to upgrade the capacity from 5 to 17 tonnes(47). This
was the first bridge in the UK over a railway strengthened
using carbon fibre plates. As the electrical supply is third
rail there were no concerns about the electrical conductivity
of the material (see Section 3.2.2).

The ends of 64 beams of the John Hart Bridge in Prince


George, British Columbia(64) were strengthened with diagonal
sheets of carbon fibre, to increase the shear capacity by about
20%. In addition to the areas that required strengthening,
carbon fibre sheet was applied to non-critical locations.
These may be removed at a later date to determine the longterm performance.
Repair following damage to the structure
The edge of the slab of the Devonshire Place Bridge in
Skipton, Yorkshire was repaired with carbon fibre plate following damage to one of the tendons(32,42). The edge beam of
a bridge in Crawley, West Sussex that had been struck by a
vehicle, was strengthened with carbon fabric.
The soffits of some of the beams of the Ibach Bridge, near
Lucerne in Switzerland, were repaired with carbon FRP plates
following damage to a prestressing tendon(65). Similarly,
repair work to the soffits of beams has been carried out in
Italy(66) to repair the damage caused by vehicle impact, the
carbon plates being used to provide some additional shear
capacity as well as increasing the flexural capacity. A beam
on Interstate Highway 95 at West Palm Beach, Florida, was
also strengthened using carbon fibre sheet after it was struck
by a truck, causing twisting and longitudinal cracking.
Insufficient reinforcement

Figure 13: Strengthening Glade Bridge.

The soffit of the Parkhouse Bridge, Helhoughton, Norfolk


was strengthened using NSM reinforcement. Two sizes of
carbon FRP bars were used, namely 16mm and 20mm, both
with a peel ply finish. The reason for using NSM rather than
carbon FRP plates was because of concern that material
floating in the river might hit the plates and remove them.
This would appear to be one of the first situations in which
NSM has been installed overhead.
Woven carbon fibre mats have been bonded directly to the
soffit of a bridge over the A10 motorway in France to
strengthen it(63). This appears to be the first application of
carbon fibre mats in Europe.
In Canada, carbon fibre sheets were applied to the soffits and
sides of the Clearwater Creek Bridge near Edmonton, Alberta,
to improve the shear resistance(14). This is a three-span highway bridge with a length of about 18m. Four beams of the
Maryland Street Bridge in Winnipeg(64) were also strengthened with vertical and horizontal sheets of carbon fibre to
increase the shear capacity by 36%. The alternative would
have been to remove parts of the bridge deck, install posttensioned external shear stirrups and cast additional concrete
round the beams. The work was carried out without interrupting the traffic on the bridge and was estimated to cost
about 70% of the conventional approach. This comparison
was based on direct costs and did not consider factors such
as traffic delays.

A bridge to the north of Wilmington, Delaware, USA had


developed longitudinal cracks because of insufficient transverse reinforcement in the bottom of the precast box-beams.
They were repaired with carbon fibre sheet(67).
Incorrect reinforcement detailing
The top surface of Haversham Bridge in Milton Keynes was
strengthened using carbon fibre plates to increase the hogging capacity. (Figure 14 shows a similar strengthening job
in Switzerland.) The plates were provided because the top
steel had insufficient lap lengths and anchorage for the
increased loading requirements. Carbon fibre plates were
chosen in preference to steel plates because of the improved
durability and the absence of the bolts required with
steel(32,34,39). (The composites are protected by the running
surface during normal operation but there is some concern
that they may be susceptible to damage when the surface is
planed off prior to resurfacing.)
The soffit of the Barnes Bridge, which carries the A34 over
the M60 Manchester Outer Ring Road, was found to have
inadequate laps in the reinforcement during an assessment of
its ability to carry 40-tonne vehicles(49). It was strengthened
with carbon FRP plate of three different sizes (Figure 15).

19

Design guidance for strengthening concrete structures using fibre composite materials

4.3.2 Columns
In this Section, the applications are grouped according to the
type of strengthening material used. The materials are
generally applied by hand (see Figure 15), though specialist
machines have been developed for large structures. These are
clamped around the column and a carrier head revolves
around the column, laying down a continuous fibre tape under
tension. The machine is gradually raised round the column, as
the required thickness of fibre is installed. Figure 16 shows
one such machine.
Wrapping with fabrics
Figure 14: Strengthening the top surface of a bridge using
carbon fibre plates.

The first trial application of FRP for the wrapping of


columns was carried out on the Bible Christian Bridge over
the A30 Bodmin Bypass in Cornwall(44,68) (see Figure 17).
Three different systems were applied to the 6m-high
800mm-diameter columns. The materials were glass, carbon
and aramid, in either sheet or ribbon form. The concrete surface was first cleaned and repaired, then generally impregnated with a thixotropic epoxy resin before the application of
the first layer of fibre. In each case, several layers of the fabric
were applied vertically, to increase the flexural capacity, as
well as in the hoop direction to increase the shear capacity.

Figure 15: Column wrapping.

Figure 17: Bible Christian Bridge, Cornwall.

Columns of the approach spans of the A92 Tay Road Bridge,


which was constructed in the mid-1960s, were wrapped with
aramid fibre sheet to improve their vehicle impact resistance(55). Similar work was carried out on the Patchway
Viaduct on the A38(56). As part of the installation, addi-tional
bands of aramid were installed above the area to be
strengthened that included deliberate defects. Trials were
subsequently carried out to assess the effectiveness of thermography in locating the defects.

Figure 16: Machine for wrapping columns.

20

Carbon sheet for wrapping columns was developed in Japan


and has been widely used for strengthening bridges, particularly to improve their seismic resistance(69). The approach
has been approved by the California Department of Transportation since the early 1990s(70).

Review of applications

In New York State, the piers of a railway bridge over a major


highway were wrapped with a water-cured prepreg glass
fabric(71). This appears to be the first use of a water-cured
material. It is not clear whether the wrapping provided additional strength or was mainly to protect the concrete, which
had suffered severe damage; the bridge was described as
structurally sound.
In Canada, repairs were carried out in August 1996 at Sainttienne-de-Bolton, Qubec, where nine columns of a bridge
over Highway 10 were repaired, five with glass fibre and
four with carbon fibre, supplied by three different companies(72,73). The circular columns are 6m high, with a diameter of 760mm. The work was backed up by laboratory
studies, including the behaviour of the wrapping materials
under wetdry and freezethaw cycles.
In Montreal, one of the main piers of the Champlain Bridge
over the St Lawrence River was wrapped in October 1996. A
total of nine layers of glass fibre wrap were installed to give
a thickness of 10mm, both to strengthen and protect the
concrete from ice damage. The column is 1.37m in diameter.
To reduce the problems associated with working over water,
the fibre sheets were preimpregnated with resin and wrapped
round a roller on land. They were then installed on the pier
by simply rolling out while the resin was still wet(73,74). This
technique should not be confused with the use of a prepreg
part-cured material.
Combined plates and wrapping
The reinforcement in circular columns of a bridge in Poland
was heavily corroded. After repair of the concrete, the area
of longitudinal reinforcement was found to be insufficient
and the links needed to be reinstated. The repaired columns
were strengthened longitudinally with carbon FRP plates and
then wrapped with carbon sheet(75).

below water. Possibly the largest application to date has been


the Yolo Causeway, west of Sacramento, California, where
3000 columns were wrapped with glass fibre reinforced
preformed shells(78). The first application in the UK was on
the A19 on Tyneside, where 24 columns were strengthened
with glass FRP shells(51,52). The original columns were
tapered, but the shells were of uniform diameter throughout.
Thus the thickness of the annulus filled with grout increased
from the bottom of the column to the top.
4.3.3 Continuity
There has been limited use of composites to improve the
continuity of bridges. In 1986 the joints in the Kattenbusch
Bridge in Germany were strengthened by bonding a large
number of glass fibre reinforced polymer composite plates
across them(65). The plates were 3.2m long, 150mm wide and
30mm thick.

4.4 OTHER STRUCTURES


4.4.1 Towers and chimneys
In Japan, deteriorated concrete chimneys have been strengthened by means of carbon or aramid fibre tapes bonded to the
surface, generally to increase the seismic resistance but also
to increase the resistance to wind and thermal loading(79).
When a former cement plant in San Antonio, Texas was
converted into a retail and entertainment complex, the chimneys were wrapped in glass FRP to increase their flexural
and shear capacities and to improve their appearance(80).
Trials are planned for NSM reinforcement strengthening on
the chimneys of a disused power station in London. In
addition, a cathodic protection system will be installed to
protect the steel reinforcement. Hence aramid FRP rods have
been selected for the NSM because they are non-conducting.

Expansive grout combined with fabric

4.4.2 Tunnels

As part of the repair of the Leslie Street bridge in Toronto(74),


an expansive mortar was cast round a deteriorated column.
The repair was wrapped with a plastic sheet, followed by a
glass fibre wrap. As the mortar continued to expand it tensioned the glass fibre, putting the parent concrete into biaxial
compression.

Carbon fibre sheets have been used in a number of highway


and railway tunnels, to repair cracks in concrete linings and
also to increase the strength. Fukuyama et al.(81) reported that
there were approximately 25 such applications in Japan in 1996.

Preformed shells
Various types of prefabricated glass fibre composite shell are
being developed in the USA, including the full height Hardcore system, as used on the Santa Monica Freeway in Los
Angeles and the segmental Clockspring system(76).
Preformed shells were used to strengthen columns on the
New Jersey Turnpike, which had heights between 3 and
4.5m(77). The shells were installed with a clearance of
50150mm from the concrete surface, which was later filled
with grout. In some locations the lower end of the shell was

A large-diameter water chamber forming part of the


Frontenac Hydroelectric Power Plant in Sherbrooke, Quebec,
Canada was strengthened in 1998 with glass FRP on both the
inside and the outside faces. This was an environment with
very high humidity and the strengthening was made more
complicated by water seeping through the highly porous
concrete(82).
4.4.3 Marine/coastal structures
Various lighthouses in the North Sea have been strengthened
with carbon fibre sheet material. The alternative, steel bands,
lifted into position using a helicopter, would have been a
more expensive option.

21

Design guidance for strengthening concrete structures using fibre composite materials

The deck of the 29-span Langstone Bridge, which connects


Hayling Island near Portsmouth to the mainland, was
strengthened with carbon fibre plates to enable to bridge to
carry 40-tonne vehicles.

Francisco, 245 submerged reinforced concrete foundation


piers were retrofitted for seismic confinement using a
combination of systems.
4.4.4 Miscellaneous structures

Following the construction of a new pier at the Humber Sea


Terminal, part of the approach span required strengthening.
Carbon fibre plates were bonded to the underside of the
approach ramp(83).
The US Navy is carrying out trials on various composite
materials for strengthening concrete piers(84). Gee(85) reports
that piles in the tidal zone were wrapped using an epoxy
specially formulated for use underwater and the strengthened
area then protected with a layer of plastic sheet until the resin
had fully cured. The supporting columns of various piers and
other coastal structures in California have been strengthened
by wrapping with carbon fibre. In some cases the strengthening extended below ground level. At Fort Mason in San

22

Vertical and horizontal bands of aramid FRP were used to


strengthen the cooling towers of West Burton Power Station
in Nottinghamshire(86). Aramid was chosen because of its
abrasion resistance. The turbine support units at Torness
Power Station were strengthened using carbon fibre sheet.
In Japan, concrete electricity transmission poles have been
strengthened using carbon fibre sheet material. In Montreal,
Canada, laboratory trials have been carried out on railway
sleepers strengthened with polyester fabric(87). A 30-year-old
processing tower in Qatar was strengthened with 3500m of
carbon FRP plate. Several options were considered but the
material was chosen because of the speed of installation(15).

STRUCTURAL DESIGN OF
STRENGTHENED MEMBERS

5.1 SYMBOLS

fcu

The following symbols are used in this Report. They are


largely compatible with BS 8110(88). It should be noted that
some of these symbols may differ from those used in Eurocode 2, EN 1992 Design of concrete structures(89).

ff
ffd
ffk
ffm
fr
fy
G
gs
h
Ics

Ae
Af
Afa
Afs
Ag
Aol
Asa

AsaF
a
b
ba
bbarperim
bf
bnotchperim
c
D
d
df
Ec
Efd
Efk
E0
E2
Es
e
fat
fcc
fccd
fc0
fctm

effectively confined area of concrete


area of FRP
additional longitudinal FRP strengthening area
due to shear
area of FRP shear reinforcement
gross cross-sectional area of section
area of overlapping parabolas for rectangular
columns
area of effectively anchored additional
longitudinal tensile steel for shear requirement in
BS 5400
increase in FRP required if Asa steel is ignored
major dimension of elliptical column
width of section
width of adhesive layer
perimeter of NSM FRP bar
width of laminate
effective perimeter of NSM notch
minor dimension of elliptical column
diameter of circular column
effective depth of section
effective depth of FRP shear reinforcement
initial modulus of elasticity of concrete
design elastic modulus of FRP
characteristic elastic modulus of FRP
secant modulus of concrete = 0.67fcu/(mc c0)
slope of linear portion of confined concrete stress
strain curve
modulus of elasticity of steel
eccentricity of load on column = M/N
design adhesive tensile strength
confined concrete axial compressive stress
design confined concrete compressive strength
unconfined concrete compressive strength =
0.67fcu/mc
tensile strength of concrete = 0.18(fcu)2/3 (ideally
derived from in situ pull-off tests)

k
kb
lol
lnsm
lnsm,max
lt
lt,max
M
Madd
N
n
ne
Q
Rc
s
sf
Tk
Tk,max
Tnsm
Tnsm,ad
Tnsm,max
tf
V
Vadd
Vc
Vf
VR,max
Vs
Vu

characteristic compressive cube strength of


concrete
stress in axial FRP at location of shear force
design tensile strength of FRP
characteristic tensile strength of FRP
mean tensile strength of FRP
confinement pressure
characteristic tensile strength of steel reinforcement
dead load
shape factor for non-circular columns
overall depth of member
second moment of area of strengthened concrete
equivalent transformed cracked section
confinement effectiveness factor
bond force factor, defined in Equation 16
length of overlapping region of parabolas
anchorage length provided for NSM bar
anchorage length for NSM bar required to generate
Tnsm,max
anchorage length
maximum anchorage length
design ultimate moment
additional required moment capacity
ultimate axial load on column
factor for anchorage of shear strengthening
number of effective axial reinforcing bars
live load
corner radius of rectangular column
standard deviation
spacing of FRP strips
characteristic bond failure force
ultimate bond failure force
characteristic anchorage force for NSM
characteristic adhesive bond failure force
maximum NSM anchorage force
thickness of FRP laminate
shear force due to ultimate loads
shear force additional to that present at the time of
strengthening
shear resistance of concrete
shear resistance of FRP
maximum allowable shear resistance of member
shear resistance of steel reinforcement
ultimate shear capacity

23

Design guidance for strengthening concrete structures using fibre composite materials

vmax
x
z

maximum permissible shear stress


depth of neutral axis of FRP-strengthened member
lever arm

e
f

modular ratio of steel to concrete


modular ratio of FRP to concrete
Angle between the principal fibres of the FRP and
a line perpendicular to the longitudinal axis of the
member
change in force over length y of FRP for longitudinal shear stress
short length along FRP for longitudinal shear
stress
partial safety factor for adhesive
partial safety factor for concrete
partial safety factor for modulus of elasticity of
FRP
design partial safety factor for modulus of
elasticity of FRP
design partial safety factor for strength of FRP
partial safety factor for manufacture of FRP
partial safety factor for steel
design partial safety factor for strain of FRP
partial safety factor for strain of FRP
axial strain in unconfined concrete at peak stress
=2.4 x 104(fcu/mc)
confined concrete axial strain
confined concrete ultimate axial strain
design ultimate strain of FRP
effective FRP strain
final strain of FRP for flexural strengthening
characteristic failure strain of FRP
effective strain in the FRP for shear strengthening
position of transition region between parabola and
straight line for confined concrete
yield strain of steel = 0.002
creep coefficient
longitudinal shear stress

ff
y

mA
mc
E
mE
mf
mm
ms
me

c0
cc
ccu
fd
fe
ff
fk
fse
t
y

5.2 OVERVIEW OF AVAILABLE DESIGN


GUIDANCE
Since the publication of the First Edition of this Technical
Report, a number of national and international guidelines
have been introduced dealing specifically with the design of
externally strengthened concrete structures. In particular the
Federation Internationale du Beton (FIB) task group 9.3 have
published Bulletin 14(90) and the American Concrete Institute
has published ACI 440.2R(2). Other guidelines have been
developed by the Japan Society of Civil Engineers(91), the ISIS
Canada Research Network(3) and by Tljsten in Sweden(92).
In the UK other relevant publications are provided by the

24

Highways Agency. Their design guide BD 84/02(10) provides


advice on strengthening concrete bridge supports using FRP.
The Agency is currently drafting further guidance on using
FRPs for strengthening highway structures. The Construction
Industry Research and Information Association (CIRIA) has
published a report on the use of composites in construction(93)
and guidelines on strengthening metallic structures using
FRPs(6). Advice on the design of adhesively bonded joints,
for fibre composite materials, is given in the EUROCOMP
design code and handbook(94).

5.3 BASIS OF DESIGN


This part of the Report provides the necessary guidance for
engineers to carry out the design of non-prestressed FRP
strengthening systems for concrete structures. It should be
read in conjunction with BS 8110(88) and BS 5400: Part 4(95)
as appropriate. It is important, however, to recognise that the
basis of strengthening using FRPs differs from the design of
conventional steel-reinforced concrete structures in a number
of important respects. These include the elastic-brittle
behaviour of FRP materials and the bond behaviour of
externally applied FRPs.
The strengthening will generally be carried out following a
detailed appraisal. When the structure is a bridge, reference
should be made to the Highways Agencys BD 44 The assessment of concrete highway bridges and structures(96), which
gives modified forms of the equations in BS 5400 for appraisal.
The partial safety factors used are lower than those used in
BS 5400, reflecting the reduced level of uncertainty, and the
use of the worst credible strength is permitted. For buildings,
the Institution of Structural Engineers Appraisal of existing
structures(8) also suggests lower partial safety factors, but
appropriate equations have not been developed, though the
approaches adopted by BD 44 should be equally applicable
to other types of structure. Where appropriate, guidance in
the present document is generally based on BS 8110.
The design of FRP strengthening systems should be based on
limit state principles. The aim of limit state design is the
achievement of an acceptable probability that the structure
being strengthened will perform satisfactorily during its
design life. This involves checking that the structure does
not reach a limit state during its intended life, which may
render it unfit for use.
Limit states broadly fall into two categories: ultimate and
serviceability. Ultimate limit states normally encompass
mechanisms that cause partial or complete collapse of the
structure while serviceability limit states correspond to states
that principally affect the appearance or proper performance
of the structure. Examples of ultimate and serviceability
limit states relevant to FRP strengthening systems are given
in Table 4.

Structural design of strengthened members

Table 4: Limit states relevant to FRP strengthening systems.

Ultimate
Strength
Bending
Shear
Compression
Anchorage/plate separation
Fire

Serviceability
Deflection
Cracking
Steel stress
Fatigue
Creep
Stress rupture
Durability

The design of FRP strengthening systems is mainly concentrated on the ultimate limit state of strength (see Chapters
6 to 8). This includes checks for bending, shear and compression, conditions normally associated with reinforced concrete
design, as well as checks for plate separation that are peculiar
to FRP-strengthened structures. Since structural strengthening
invariably increases the stiffness of flexural members, which
in turn increases the risk of brittle failure, a check on ductility
will also be necessary (see Section 6.2.4). Since the proportional increase in stiffness will be less than the increase in
strength, it will also be necessary to check the deflection of
the strengthened structure against the appropriate limits.
Service loads should not adversely affect the appearance or
efficiency of strengthened structures. Generally, FRPstrengthened structures should experience closely spaced
narrow cracks provided that good bond exists between the
FRP and the concrete substrate. However, where problems
are anticipated the designer should take steps to ensure that
the design crack widths do not exceed the limits recommended in BS 8110 or BS 5400, as appropriate. The steel
reinforcement should not yield under the service load,
otherwise permanent deformations in the structure will result.
Fatigue and stress rupture are taken into account by using
lower design stresses determined in accordance with Section
6.8. Much of the testing work that has been carried out has
confirmed that carbon FRP retains its chemical and physical
properties when exposed to conditions typical of those
relevant to concrete construction. However, other materials
are less stable when exposed to moisture or ultraviolet
radiation, and consideration must therefore be given to the
use of protective coating systems (see also Section 10.11).
For buildings, fire should also be included in the above limit
states as it will influence the properties of both the FRP and
the adhesive used to attach the FRP to the concrete (although
some fire-rated structural systems are available). This aspect
is discussed further in Section 5.7.1.

5.4 MECHANICAL PROPERTIES OF


MATERIALS
5.4.1 Properties of concrete and steel reinforcement
The strength of the concrete to be used in the design equations
given in Sections 6 to 8 should be the characteristic (28-day)
compressive cube strength, fcu. The characteristic tensile
strength of modern mild steel and high-yield steel reinforce-

ment, fy, should be taken as 250 and 460N/mm2, respectively.


Both steel types have a mean modulus of elasticity, Es, of
200kN/mm2. Different steel strengths may be appropriate for
historic structures.
Where there is sufficient knowledge of the properties of the
actual materials in the structure, modified values may be used.
BD 44(96) uses the concept of the worst credible strength,
both for the steel and the concrete. Where actual values are
available, modified values for the partial safety factors given
in Section 5.6 may be used.
5.4.2 Properties of fibre-reinforced polymers (FRP)
The mechanical properties of FRP materials depend principally on the type and percentage of fibre used. These aspects
are likely to vary between competing composite products,
and since there is currently no agreed standard specification
for their manufacture, all design must be on the basis of the
actual properties obtained from the manufacturer, who should
supply either characteristic values or mean values and standard deviations. Only materials manufactured in accordance
with an approved quality control scheme should be used;
Appendix C gives guidance on the appropriate level of
testing required.
The normally available mechanical properties of FRP are
tensile strength, modulus of elasticity and elongation at
failure. For fabric materials, the mechanical properties may
be measured directly on samples, which may be assumed to
be representative of the material applied to the structure. The
properties of plates should be determined on representative
samples. For wet lay-up systems, test samples should be
prepared under the same conditions as the composite is
applied to the concrete. Fully cured samples may then be
tested to give an indication of the in situ properties. The
mechanical properties of other manufactured composites,
such as shells, should be determined by the manufacturer
from tests on coupons.
For example, the characteristic tensile strength of FRP, ffk, is
related to the mean tensile strength, ffm:
ffk = ffm 2 s

(Equation 1)

where s is the standard deviation. (Note: sufficient samples


must be tested to ensure that two standard deviations is
realistic, see Appendix C.)
Tables B2 and B3 in Appendix B give typical mechanical
properties for a range of FRP strengthening systems that are
currently available. The information is taken from manufacturers data sheets and is thought to be correct at the time
of publication. For design purposes, actual properties must
be obtained from the manufacturer. As test methods vary, the
information should detail the basis for the information (e.g.
frequency of testing, standard deviation).

25

Design guidance for strengthening concrete structures using fibre composite materials

5.4.3 Properties of adhesives and laminating resins


Tables B5 and B6 in Appendix B give typical properties for
epoxy adhesives and laminating resins that are currently
available. (The comments in Section 5.4.2 regarding the
information on FRP are equally applicable.) It is important
that the adhesive or laminating resin being used is compatible
with the laminate or fibre. Ideally, to ensure compatibility,
all the components of the system (including any priming or
top coating materials) should be from a single supplier.
5.4.4 Stressstrain curves
The equations developed for the design of FRP strengthening systems given in Sections 6 to 8 are based on the
rectangular parabolic stressstrain relationship for concrete
in compression and the relationship for reinforcing steel
described in Section 6.2.4, which gives the reason for
adopting this relationship. Alternatively, more exact stress
strain curves, such as those given in Eurocode 2(89), may be
used in analysis.
Unlike steel reinforcement, all FRP has a linear elastic
response to failure, with no or very limited yielding. Woven
fabrics have a degree of non-linearity, but this may be
ignored for design purposes.

5.5 PARTIAL SAFETY FACTORS FOR


LOADS

The partial safety factors are intended to take into account


the uncertainties associated with the material itself and with
its use in the structure. Guidance on developing projectspecific partial safety factors can be found in the CIRIA
report on FRPs in construction(93). However, in most situations and in the absence of independent field-testing of material properties as installed, the partial safety factors given in
the following Sections may be used.
The magnitude of the partial safety factors applied to the
FRP will depend on the type of fibre and the stage in the
manufacturing process at which the test samples are taken
(see Section 5.6.3). The partial safety factors are intended to
take into account changes in material properties with time. In
this respect they differ from the factors applied to traditional
construction materials such as steel and concrete, whose
properties are assumed not to change with time.
5.6.2 Design strength of steel and concrete
In general, the design strength of the steel and concrete
should be assessed using appropriate values of the partial
safety factors for concrete, mc, and for steel reinforcement,
ms, given in BS 8110 or BS 5400 as appropriate. When the
worst credible strengths are used (see Section 5.4.1), the
factors may be modified in line with the recommendations of
BD 44(96).
5.6.3 Design elastic modulus of FRP

For the ultimate and serviceability limit states, the design


loading will normally be obtained by multiplying the characteristic dead and imposed loads by appropriate partial safety
factors taken from BS 8110 for buildings or from BS 5400
or BD 37(97) for bridges.
Prior to strengthening, the designer will need to assess the
probable effect of an accidental loss of strengthening effectiveness resulting from fire, vandalism or impact. Guidance
on assessing the flexural strength of structures in fire is given
in Section Four of BS 8110: Part 2. See also Section 5.7 of
this Report.
If the magnitude of the initial strains that exist in the structure need to be estimated in order that they can be excluded
from the strain in the FRP (see Section 6.2.5, part (c)), the
partial safety factor for dead load and imposed load should
be taken as 1.0. The level of imposed load to be used in the
calculations will be very structure-dependent and will be a
matter for judgement.

5.6 DESIGN VALUES FOR MATERIAL


PROPERTIES
5.6.1 Introduction
The characteristic material properties (see Section 5.4) are
divided by appropriate partial safety factors (mE, m, mm)
from Tables 5 to 7 to give the values to be used in design.

26

In most practical design situations the limiting factor governing the failure of an FRP-strengthened structure is the strain
in the FRP (e.g. anchorage, separation failure, etc.), although
rarely ultimate strain. It is therefore the stiffness of the FRP
that is of importance. Although durability tests in laboratory
conditions on unloaded glass and carbon FRP composites
have shown that there is little significant degradation of the
modulus of elasticity under long-term (10,000hr) environmental exposures such as salt water, high alkalinity, humidity
and freezethaw(98), the modulus of elasticity of FRP may
change with time under load and may vary according to the
method of manufacture and application. In particular, lack of
straightness of fibres can significantly affect the stiffness. In
addition, the accuracy with which the properties are obtained
from test samples is dependent upon the method of manufacture. Therefore, it is necessary to apply partial safety
factors relating to both material type and method of manufacture to the modulus of elasticity of FRP in arriving at the
design strength of structures strengthened with external
reinforcement:
Efd = Ef / mE

(Equation 2)

where

mE = E mm

(Equation 3)

Structural design of strengthened members

Recommended partial safety factors for modulus of elasticity are given in Table 5 and partial safety factors for
method of manufacture and application can be taken from
Table 6.
Table 5: Partial safety factors for Youngs modulus at the ultimate
limit state.

Material
Carbon FRP
Aramid FRP
AR glass FRP
E glass FRP

Table 6: Recommended values of additional partial safety factors,


to be applied to manufactured composites, based on Clarke(94).

Additional partial
safety factor, mm

1.05
1.1
1.2

In some instances the ultimate tensile strength of the FRP is


required during the design of a strengthened structure (e.g.
ultimate flexural strength, although actual failure is likely to
be due to separation of the FRP from the concrete, which, in
general is related to FRP strain). In such cases, the design
strength can be derived from the design modulus of elasticity, Efd, and the design strain, fd:
ffd = Efd fd

(Equation 6)

mf = mE m = E (mm)2

(Equation 7)

So the design strength is given by:

1.05
1.1
1.2
1.5

ffd = ff / mf

It is also possible in some situations that the ultimate strain


in the FRP may govern failure of a strengthened structure
(e.g. shear strengthening or ultimate strain of confined concrete) although, typically, other strain limits are reached first.
Again, durability tests on unloaded specimens of glass and
carbon composites have demonstrated significant long-term
ultimate strain reductions, particularly due to exposure to
humidity(98). As for the material partial safety factor for
modulus of elasticity, the partial safety factor for ultimate
strain is also related to both material type and route of manufacture and application. Thus, the design strain is given by:

(Equation 8)

It should be noted that the resulting material partial safety


factor for ultimate strength is equivalent to that proposed in
the First Edition of TR55. Figure 18 shows the relationship
between design stress and strain and the partial material
safety factors.
Stress

ffk

Ef

ffd
Ef

(Equation 4)

Strain

Figure 18: Assumed stressstrain curves.

where

m = mm

Partial safety factor,


1.25
1.35
1.85
1.95

Hence, given the partial safety factors acting on the modulus


of elasticity and the ultimate strain of the FRP, the partial
safety factors acting on the strength of the material are as
follows:

1.05
1.05
1.1

5.6.4 Design ultimate strain of FRP

fd = fk / m

Material
Carbon FRP
Aramid FRP
AR glass FRP
E glass FRP

5.6.5 Design ultimate strength of FRP

Factor of safety, E
1.1
1.1
1.6
1.8

Type of system (and method of


application or manufacture)
Plates
Pultruded
Prepreg
Preformed
Sheets or tapes
Machine-controlled application
Vacuum infusion
Wet lay-up
Prefabricated (factory-made) shells
Filament winding
Resin transfer moulding
Hand lay-up
Hand-held spray application

Table 7: Partial safety factor for strain at the ultimate limit state.

(Equation 5)

Recommended partial safety factors for ultimate strain are


given in Table 7 and partial safety factors for manufacture
method can, again, be taken from Table 6.

5.6.6 Steel stress


The designer must check that the steel reinforcement does
not yield under service loads, otherwise the structure may
sustain permanent deformations. Therefore it is recommended
that the partial safety factors for steel reinforcement be
increased to 1.25 in performing this check. This condition
may be rather onerous for older structures reinforced with

27

Design guidance for strengthening concrete structures using fibre composite materials

grade 230 steel, which would otherwise be considered suitable for strengthening. Under these circumstances, designers
may consider increasing the allowable steel stress to 1.0 fy,
provided that other factors, e.g. crack widths and concrete
quality, do not preclude this approach to strengthening.
5.6.7 Deflection and cracking
The deflections and crack widths in structures being strengthened should be kept within values specified in current codes
and standards, such as BS 8110 or BS 5400. The calculations
should be based on the partial safety factors contained in
these documents.
5.6.8 Adhesive
In general, the ultimate behaviour of a strengthened section
will be governed by the strength of the concrete and not by
the strength of the adhesive, provided the following are
satisfied:

All the materials used are in accordance with recognised


standards.
The material properties are checked on samples made on
site.
The in-service temperature does not differ significantly
from that at which the test samples were made and cured.
The work is carried out by suitably experienced staff, in
accordance with the advice in Chapter 10.
Detailed and proven method statements and specifications are used.
The structure is fail-safe, i.e. failure of the strengthening will not lead to failure of the structure.

If any of the above parameters are not satisfied, higher values


of mA, the partial safety factor for adhesive, will be required.
An approach for determining appropriate partial safety factors
may be found in A guide to the structural use of adhesives(24).
It should be noted that cyclic strains applied to an adhesive
during the curing period, for example due to traffic loading
on a bridge under repair, may lead to a change in the
properties of the adhesive. However, it has been suggested
that these changes are likely to be small, perhaps a 10%
reduction in the strength of the fully cured material.
As a general recommendation, the sustained stress in the
adhesive should be kept below 25% of the short-term strength,
which equates to the recommended minimum material partial
safety factor of 4.0.

5.7 EXTREME LOADINGS


5.7.1 Behaviour of structures in fire
For design of reinforced or prestressed members, fire
resistance is generally ensured by the provision of adequate
cover to the reinforcement. More detailed analysis can be
carried out working from a standard timetemperature curve,

28

such as that given in BS 476: Part 20(99), for the required fire
endurance. A design approach, only for members in flexure,
is given in Section Four of BS 8110: Part 2, which gives
reduced values for the strengths of the steel and concrete at
elevated temperatures. However, as fire is considered as an
accidental load, the partial safety factors on the materials are
reduced and, more importantly, the partial safety factors on
the dead and live loads are also reduced. Thus in many cases,
the fibre composite strengthening could fail completely without risking failure of the structure. This may be illustrated by
the following simple example, which ignores any material
property changes due to the elevated temperatures:
1. Consider a floor slab with a dead load G and live load Q
2. Original design capacity = 1.4 G + 1.6 Q
3. The slab is strengthened to carry an additional 50% live
load
4. Hence modified design capacity = 1.4 G + 1.6 (1.5 Q)
5. Required design capacity in fire = 1.05 G + 1.0 (1.5 Q)
6. In a typical floor slab, G Q, so required design capacity
in fire = 2.55 G
7. Capacity of unstrengthened slab = 3.0 G which is greater
than the required capacity in fire.
If failure of the fibre composite strengthening in fire would
lead to the collapse of the structure, it will obviously be
necessary to consider the behaviour of the fibre composite
materials as well as the behaviour of the adhesive. The fibres
themselves are unlikely to be affected by the elevated temperature. However, it is likely that the adhesive will be affected
if the temperature exceeds the adhesives glass transition
temperature, which may be of the order of 50 to 60C for
conventional materials. If the glass transition temperature is
reached then the effectiveness of the FRP strengthening will
be reduced.
Specific advice on fire resistance of FRP materials should be
sought from the manufacturer. If necessary, options for increasing the fire resistance of FRPs may include providing a
layer of suitable insulating material over the fibre composite.
The carbon fibre mats used to strengthen the soffit of a bridge
over the A10 motorway in France were covered with a layer
of plaster and mortar for fire protection(100). Unless a rigorous
analysis is undertaken it is sensible to neglect the strengthening from FRP in fire situations. As shown above, such a
situation does not preclude the use of FRP strengthening.
In addition to concerns about the structural behaviour of
strengthened structures in fire, the emission of smoke and
toxic fumes will be a major consideration, particularly in
enclosed situations such as tunnels. Improved performance
can be achieved by the use of appropriate fillers and the use
of intumescent coatings or other high-temperature foam
insulation barriers.
5.7.2 Seismic loading
Seismic loading will not be a major loading case for most
UK structures. However, it may be important for strengthening work in connection with nuclear-related structures.

Structural design of strengthened members

This is a highly specialised area of design, which is outside


the scope of this Technical Report. Reference should be
made to publications by experts in the field, such as Priestley
et al.(101), Seible et al.(102) and Triantafillou(103).
5.7.3 Impact loading
The consequences of structural collapse due to vehicle impact
on bridge supports are considerable. In an impact, about 80%
of the energy is absorbed by the vehicle crushing, with the
remainder being absorbed by the structural element. It is
necessary to ensure that both the flexural strength and
deformability of a column are adequate. The designer should
ensure that the deformability of the strengthened structure is
at least as great as it would be for the equivalent conventionally designed structure.

While both issues are clearly of vital importance to security,


public-domain blast test results are notoriously difficult to
come by. Furthermore, much of the research conducted on
FRP strengthening of structures against blast has been undertaken on concrete masonry walls under static loading(105).
However, some studies of the effects of real blast loading on
FRP-strengthening schemes against debris projectiles have
been carried out(106).
It seems that masonry walls can indeed be strengthened to
resist disintegration during blast loading(107). However, an
obvious question that remains is what effect the increase in
strength might have on an increase in stiffness of the same
wall. Such an effect could be problematic in seismic
zones(108).
5.7.5 Vandalism

It has been shown that there is an increase in concrete


strength at very high loading rates. While it is not clear
whether there is such an increase in concrete strength under
vehicle impacts, tests commissioned by the Highways Agency
on circular columns wrapped with aramid FRP have shown
that the wrapping is at least as effective under impact loading
as under static loading. Therefore, it would seem that methods
of designing for impact loads based upon the application of
equivalent static loads are reasonable.
Tests on rectangular columns carried out by Suter et al.(104)
have shown the effectiveness of longitudinal aramid FRP
followed by hoop wrapping in increasing the flexural capacity and hence the energy-absorbing capacity of columns
under equivalent static loading. It should be noted that tests
have only been carried out using aramid FRP, due to its
toughness. However, this does not mean that other FRP types
are necessarily unsuitable.
5.7.4 Blast loading
There are two fundamental issues for the designer to
consider when blast is involved. The first is strengthening of
the structure to withstand the blast so that the structure is
serviceable immediately following overload. The second is
providing protection to the public by preventing the faade,
in particular, from disintegrating into projectile debris.

In situations where deliberate vandalism is considered to be


a potential problem (primarily in an urban environment)
there are a number of possible actions that can be taken.
Firstly, the FRP may be physically protected, by providing
some form of barrier that limits accessibility (either to the
surface of the strengthened structure or around the structure
as a whole). Secondly a form of strengthening which is inherently resistant to physical attack, such as NSM, may be
chosen if possible. Thirdly, frequent inspection or monitoring should be carried out so that any damage resulting
from vandalism can be quickly remedied. However, it is not
possible for these measures to prevent damage from a determined attack. It is therefore necessary, as in the case of fire
damage, for the unstrengthened structure to be able to satisfactorily carry the unfactored service loads.
Accidental damage may also occur due to contractors drilling
through FRP or removing protective coatings without realising the structural implications. This can be avoided by
making contractors aware that the FRP should not be interfered with. This may be achieved by application of some
form of printed warning, either directly on to the FRP or in
close proximity if the finish of the FRP is important. Further
guidance and suggested warning signs can be found Section
10.12 and in the Concrete Societys TR57(5).

29

STRENGTHENING
MEMBERS IN FLEXURE

6.1 GENERAL
The flexural strength of reinforced concrete beams and slabs
can be increased by bonding FRP laminates to the tension
faces of the members, as shown in Figure 19. For members
strengthened in flexure the following should be considered:

maximum moment
risk of peeling failure at the ends of the FRP
risk of debonding of the FRP and the concrete substrate
shear capacity of the section
ductility of the strengthened member
compliance with relevant serviceability limit states, e.g.
cracking, deflection, fatigue, creep-rupture.

of the concrete to ensure a proper bond and long-term


durability. Tests to determine these properties (e.g. pulloff tests) should be outlined in the specification.
The surface preparation of the concrete substrate is sufficient to achieve the required level of bond strength
required in the design.

In addition, this design guidance is dependent on assumptions based on the installation methods and specification
detailed in Chapter 10.

6.2 MOMENT CAPACITY


6.2.1 Introduction
The section should be designed such that yielding of the steel
reinforcement precedes both compressive failure of the concrete and tensile failure of the FRP.
In cases where the FRP will theoretically reach its design
tensile strain before the concrete crushes, failure normally
occurs due to plate separation rather than plate rupture and
the strain limits for debonding discussed in Section 6.3.3
will frequently govern the design. In some cases the concrete
will crush before the FRP reaches its design tensile strain
(see Section 6.2.4). Provided that the steel strain at failure is
sufficiently large, however, this should not result in brittle
failure of the strengthened member.

Figure 19: Strengthening beams and slabs with FRP.

In addition, this design guidance is dependent on the following assumptions:

No slip between the FRP strengthening and the substrate


(i.e. plane sections remain plane). This assumption places
limits on adhesive thickness, adhesive shear modulus and
FRP composite in-plane shear rigidity.
Inter-laminar shear strength of the FRP strengthening is
greater than the adhesive bond shear strength. This should
be covered in the specification by specifying the type of
resins that are acceptable, limits on fibre volume fraction
and elastic modulus of the FRP strengthening.
The substrate quality is such that it will not reduce the
effectiveness of the FRP strengthening. Therefore the
actual condition must be established and taken into
account in design together with likely future deterioration which may be indicated by chloride levels, existing
cracks, moisture and half-cell potential of the concrete
substrate. The specification should also outline clearly
the allowable minimum compressive and tensile strengths

Design ultimate moments should normally be determined by


linear elastic methods. If there is evidence of local yielding
taking place, the results of an elastic analysis need to be
applied with care. Since members undergoing strengthening
will usually be steel reinforced, some redistribution of elastic
moments may occur near ultimate. Section 6.7 gives further
guidance on redistribution.
6.2.2 Requirements of the existing section
The ultimate capacity of the existing section should be
assessed by conventional concrete design methods, such as
those in BS 5400-4:1990 clause 5.3.2.1 or BS 8110-1:1997
clause 3.4.4.1 as appropriate to the structure being considered.
Particular care should also be taken to apply appropriate
material parameters and factors for the age of the structure in
question, in order to reflect likely variability in materials at
the time of their incorporation into the structure. In particular, since many of the structures that require strengthening
will have been built prior to publication of the 1997 edition

31

Design guidance for strengthening concrete structures using fibre composite materials

of BS 8110, it would seem appropriate to use a partial safety


factor for steel, ms, of 1.15 rather than the value of 1.05 now
recommended by the code.
The section should only be considered for strengthening if
the ultimate capacity of the unstrengthened (existing) section
is at least as great as the effects arising from the unfactored
loads to be applied. This ensures that even in the event of
removal of the FRP strengthening by some unforeseen event,
catastrophic collapse of the structure is not likely.

6.2.3 Preliminary design

An initial but potentially non-conservative estimate of the


FRP requirement for the section can be obtained by assuming that the position of the neutral axis remains approximately equal to that of the unstrengthened section. The
approximate area of FRP required, Af, can therefore be
obtained by dividing the required additional moment capacity of the beam, Madd, by the product of the steel lever arm,
z , and the design stress in the FRP (given by fe Efd) as follows:
Af = Madd / fe Efd z

(Equation 9)

where
fe
= the lesser of fk / m (the design ultimate strain of
FRP) and a strain of 0.008. Typically the value of
0.008, which would result in separation failure (see
Section 6.3.3) governs.
Efd = design modulus of elasticity of FRP, Efk /mE
z
= steel lever arm
This calculation becomes a less reliable predictor of the FRP
area required if the existing section is already heavily reinforced, or if the section is doubly reinforced. In any event, it
is always necessary to proceed with the detailed design
method rather than rely on this initial estimate.
6.2.4 Design resistance moment of FRP-strengthened
beam
When analysing a cross-section to determine its ultimate
moment of resistance the following assumptions should be
made:

32

The strain distribution in the concrete in compression and


the strains in the reinforcement, whether in tension or
compression, are derived from the assumption that plane
sections remain plane and that no longitudinal slip occurs
between or within the components of the section.
The stresses in the concrete in compression are derived
from the stressstrain curve in either BS 5400-4:1990
Figure 1 or BS 8110-1:1997 Figure 2.1, with the strain at
the outermost compression fibre at failure taken as no
more than 0.0035.
The tensile strength of the concrete is ignored.

The stresses in the metallic reinforcement are derived


from the stressstrain curves in BS 8110-1:1997 (see
Figure 20). This condition is more onerous and appropriate for longitudinal shear than the stressstrain curve
in BS 5400-4, while having negligible effect on ultimate
moment capacity. The material partial safety factor for
steel, ms, should be set to an appropriate value for the
date of construction and with the compressive stress
limited to fy/(ms+fy/2000), as it is in BS 5400-4.
The strains in the FRP reinforcement take into account
the strains present in the bonded surface at the time of
application of the reinforcement.
The stresses in the FRP reinforcement are derived from
the assumption that the FRP has a linear elastic characteristic until rupture.

In addition, if the ultimate moment of resistance, calculated


in accordance with this clause, is less than 1.15 times the
required value, the section should be proportioned such that
the strain at the centroid of the tensile reinforcement is not
less than 0.002 + fy/(Esms).

stress
fy/ms

E=200 kN/mm2
strain
Figure 20: Stressstrain curve for reinforcing steel in the
design of strengthened beams in flexure.

Allowance should also be made for the requirements of


additional tensile capacity (to carry the tensile forces arising
from the truss analogy for resisting shear) if significant shear
and bending moment coincide, as would be the case when
strengthening a continuous structure over a support.
6.2.5 Example design method
The requirements of Section 6.2.4 can be met by adopting
the following example design method at any section that
may be critical:
(a) Calculate the loads to be applied to the structure, determining loads applied before the strengthening (such as
element self-weight) and loads after strengthening (such
as traffic loading in the case of a bridge) separately. From
these determine both shear forces and bending moments
at the section considered. To simplify later stages these
load effects should also be determined both unfactored
and with ultimate limit state load factors applied.
(b) Calculate the area of reinforcement necessary to satisfy
the Asa requirements of BS 5400-4:1990 clause 5.3.3.2.
(c) Calculate the strain in the section at the position where
the FRP is to be applied under the unfactored loading

Strengthening members in flexure

present at the time the strengthening is to be applied. In


calculating this value, appropriate allowance should be
made for the duration of the loads, since these loads are
likely to be predominantly dead loads and consequently
the modulus of elasticity adopted should be that for longterm loading. In calculating these strains the following
assumptions should generally be adopted:
Plane sections remain plane.
The reinforcement, whether in tension or compression,
is elastic with a modulus of elasticity of 200kN/mm2.
The concrete in compression is elastic with an appropriate modulus of elasticity as discussed above. This
may be half the value in Table 3 of BS 5400-4:1995.
The concrete has zero tensile capacity.
(d) Assume an initial concrete maximum compressive strain,
which should be less than 0.0035.
(e) Assume an initial neutral axis position.
(f) Adopting the assumptions described in Section 6.2.4
calculate the forces in the component parts of the crosssection. The strain used to calculate the force in the FRP
should be evaluated by subtracting the initial strain in the
concrete at the position of the FRP at the time of
strengthening (calculated in step (b)) from the strain at
the position of the FRP from the assumed linear strain
profile (dependent on the assumed neutral axis position
and maximum concrete strain in steps (d) and (e)). Optionally, the area of existing steel tensile reinforcement
assumed in this calculation should be the provided area
minus the value of Asa required at this section, as calculated at step (b).
(g) Iteratively adjust the assumed neutral axis position and
concrete maximum compressive strain until step (f) results
in zero net axial force present in the section (i.e. force
balance is achieved) and the moment of these forces
matches (or exceeds) the required bending moment.
(h) Check the calculated stresses and strains against the
following criteria:
The concrete maximum compressive strain should
not exceed 0.0035.
The FRP maximum tensile strain should not exceed
the limit calculated in accordance with Section 6.3.3.
The section exhibits adequate plasticity. In the absence
of more rigorous examination, the strain at the centroid
of the tensile reinforcement should be not less than:
0.002 +

fy
Es ms

(Equation 10)

unless the ultimate moment that can be resisted by the


section is at least 1.15 times the applied ultimate
moments.
The stress in the FRP is less than the ultimate capacity
of the FRP.
(i) If the area Asa was not subtracted at step (f), the area of
FRP calculated shall be increased by an area AsaF, where:
AsaF = V /(2 ff Efd)

(Equation 11)

where
V
= the shear force due to ultimate loads
ff = the final strain in the FRP at step (g)
Efd = Efk / mE
If the calculation described does not converge at a solution
that meets the criteria described, then the quantity of FRP to
be applied may need adjustment and the calculation repeated
until an adequate design is achieved.

6.3 FRP SEPARATION FAILURE


6.3.1 Introduction
For members strengthened in flexure, failure can occur when
there is a loss of composite action between the FRP and the
concrete section. Typically failures occur through the development of a longitudinal failure-plane close to the interface
between the FRP and the concrete or at the level of the main
reinforcement. In experimental studies, FRP separation has
been found to be the most common failure mechanism. It is
therefore essential that it is taken into account in the design
of FRP strengthening schemes. Figure 21 illustrates typical
FRP separation failures observed in tests.

Figure 21: Possible failure modes and locations for FRPstrengthened beam.

Despite the importance of FRP separation, it remains a subject


that stimulates considerable research. A number of different
initiation mechanisms have been identified and proposals to
categorise them developed, for example see Blaschko et
al.(109) and Teng et al.(110). While the precise mechanisms are
still the subject of some debate, the design approach given
here has been used in many practical strengthening schemes
and is recommended.
It should be noted that the proposed method is largely based
on laboratory and field data obtained from strengthening
schemes using carbon FRP. It may not therefore be directly
applicable to other composite materials, although the same
principles should apply. In some cases the approach may be
unduly conservative for composite materials other than
carbon.
Work on steel plate bonding has shown that separation
failures tend to initiate from the ends of the plates. To address
this, limitations on plate aspect ratio are incorporated in the

33

Design guidance for strengthening concrete structures using fibre composite materials

Highways Agency Advice Note on steel plate bonding, BA


30(9), together with requirements for bolting. Early work on
FRP separation failures similarly focused on the ends of the
plates.
However, for FRP strengthening schemes, experimental evidence now shows that separation can also initiate from
flexural cracks in the span, shear cracks or concave irregularities in the surface profile, and that all of these cases need
to be taken into account in the design. Importantly, research
has also shown that externally bonded FRP strengthening
can be highly effective without the need for bolting or the
use of other mechanical fixings.
It has been shown that increasing the area of FRP bonded to
the concrete and reducing the FRP thickness reduce the
likelihood of separation failure modes.
6.3.2 Bond failure

for varying anchorage lengths and a range of different FRP


plate thicknesses. From this Figure it can be seen that, for all
the cases considered, the maximum force that can be developed in the FRP anchorage is less than 25% of the ultimate
FRP capacity.
Experimental studies have, however, shown that the FRP
force that can be developed in the span of strengthened beams
can be very much greater than the FRP anchorage capacity.
These findings indicate that, provided there is a gradual buildup of stress outside the anchorage region, it is possible for
the FRP to sustain stresses in excess of the anchorage capacity
without separation failure occurring. Importantly, it seems
that this gradual build-up of FRP stress relies on some
flexural cracking of the concrete as the ultimate limit state is
approached. Thus, particular care is required in cases where
the FRP is bonded to concrete that is not expected to crack
at the ultimate limit state, for example because of changes in
section properties or the presence of prestress.

The bond behaviour of externally bonded FRP differs markedly from that of embedded steel reinforcement. Experiments have shown that the longitudinal shear stress that can
be transferred between the FRP and the concrete is not
independent of the bonded length, as typically assumed for
embedded steel reinforcement. Thus, while it is possible to
anchor steel reinforcement by providing an anchorage length
beyond which the full strength of the reinforcement can be
developed, this is not typically the case for externally bonded
FRP. This aspect of the behaviour of externally bonded FRP
greatly influences, and adds complexity to, the design of
strengthening schemes.

6.3.3 Design procedure

In tests on the anchorage of FRP externally bonded to concrete, it has been found that beyond a limiting bonded length,
of the order of 50300mm, there is no further increase in the
ultimate anchorage load-capacity with increased bonded
length. Furthermore, this ultimate anchorage capacity can be
very much less than the ultimate tensile capacity of the FRP.

The presence of shear cracks can lead to a tendency for a step


to develop in the tension face of the member to which the
FRP is bonded. This can result in the development of sizeable transverse tensile stresses in the adhesive and the surface
concrete, leading to the initiation of FRP separation failure.

E frp
frp
F ctm
Kb

1.4

frpdb, max /frpu

1.2

= 230GPa
= 0.015
= 3MPa
=1

frp = frpu

1
0.8

frp = frpdb (Neubauer & Rostasy)

0.6

frp = 0.5 frpu

0.4

frp = 0.004

0.2
0
0

0.2

0.4

0.6

0.8

1.2

tfrp (mm)

The design procedure to account for FRP separation failures


first requires two structure-dependent conditions to be
checked, namely that failure will not be initiated either by shear
cracking or by irregularities in the concrete surface profile.
Provided these are satisfied, three further design-specific
criteria must be considered relating to the strain in the FRP,
the longitudinal shear stress between the FRP and the
concrete and the stresses developed in the anchorage region.
Shear-crack-induced FRP separation

Such a mode of failure may be disregarded if the maximum


applied shear force can be carried by the concrete alone,
neglecting any contribution to the shear capacity provided by
shear reinforcement. However, for beams this criterion may
well not be satisfied. Experimental studies(113) have shown
that shear cracking will have initiated at or before 67% of the
ultimate shear capacity of the section (including all forms of
reinforcement). Therefore should the maximum applied shear
force exceed 67% of the ultimate capacity, it may be presumed
that shear-crack-induced separation failure will occur. Should
the maximum applied shear force lie between the concreteonly shear capacity and 67% of full capacity, careful consideration should be given to shear crack initiation.

Figure 22: Variation in FRP separation strain with bonded


length, based on Denton et al.(112).

Surface irregularity-induced FRP separation

Such behaviour is shown in the work of Neubauer and


Rostasy(111). Based upon their model, the ratio of the strain
when FRP separation occurs to the ultimate FRP strain
capacity is plotted in Figure 22 (based on Denton et al.(112))

Concave irregularities in the profile of the surface to which


the FRP is bonded will lead to the development of tensile peeling stresses in the adhesive and surface concrete as the FRP
attempts to straighten under load. Such transverse tensile
stresses can promote the initiation of FRP separation failure.

34

Strengthening members in flexure

It is usually the case during strengthening works that the


surface to which FRP will be bonded is concavely curved to
some extent. Such unevenness is sometimes relatively local,
perhaps due to formwork being flexible during casting, or
alternatively it could also be more global, for example when
the entire soffit of a structure is curved.
Through testing(114,115), it has been found that concave curvature can significantly affect the degree of strengthening
achieved. The work of Eshwar et al.(114) suggests that the
extent over which the concave curvature exists may affect the
significance of such concavity, and it seems that the behaviour of strengthened members is more sensitive to global
than local curvature.
It is advised therefore that if the soffit of a concrete structure
to be strengthened is globally concave, reference should be
made to specialist literature(114,115), and specialist advice
sought.
If undulations in the concrete are local with a smoothly
varying profile, the influence of curvature may be disregarded in the design provided over any 1-m length, any
concavity in the FRP profile, as installed, does not exceed
3mm in depth. Fabric-based systems tend to closely follow
the profile of the concrete to which they are bonded and it is
therefore essential that the Specification for such schemes
requires the concrete surface to which the FRP will be
bonded to have only smooth variations in profile with a
maximum unevenness of 3mm in 1m. For plate-based
systems, the FRP tends not to follow the profile of the
concrete so closely and therefore greater unevenness in the
concrete profile, up to 5mm in 1m, may be acceptable
provided the FRP, once installed, has a smooth variation in
profile with a maximum unevenness of 3mm in 1m. In such
cases, the difference in the concrete and FRP profile must be
taken up in the adhesive.
Maximum FRP strain
Early work on FRP separation failures sought to establish
design values of the FRP strain below which separation would
not occur. Such design values were typically rather less than
the ultimate FRP strain capacity. Neubauer and Rostasy(111)
suggest a limit of 5y (critical for mild steel) or half the
ultimate plate strain, which for the materials tested was
0.0075. Other workers have suggested somewhat lower limits,
in the order of 0.006 for sagging moments and 0.004 for
hogging moments(116).
Such an approach alone does not fully capture the mechanics
that underpin the initiation of FRP separation failure. However, it does seem prudent at present to retain a limit on the
maximum design FRP strain. When used in conjunction with
a limit on the maximum longitudinal shear stress between the
FRP and the concrete, UK experience suggests that higher
strain limits than those described above are reasonable.

It is therefore recommended that the strain in the FRP at


ultimate limit state should nowhere exceed 0.008. This strain
limit will be more critical than the ultimate FRP strain
capacity in the substantial majority of design cases,
particularly when standard modulus materials are used. It
may be quite conservative in some cases and may potentially
be relaxed if specialist advice is sought.
Longitudinal shear stress between FRP and concrete
To ensure that the build-up of stress in the FRP outside the
anchorage region is sufficiently gradual, it is recommend
that the longitudinal shear stress between the FRP and the
concrete, determined as described below, should nowhere
exceed 0.8N/mm2 at the ultimate limit state. This value is
mentioned in the current Highways Agency Advice Note for
steel plate bonding, BA 30(9), and is broadly in line with both
the allowable shear stress values recommended in BS 8110:
Part 1 and BS 5400: Part 4. Provided specialist advice is
sought, it may be reasonable to moderately increase this
limiting longitudinal shear stress in special cases.
For surface-mounted reinforcement, when both the original
section and the applied FRP are prismatic and do not taper
along their length, assuming the concrete and steel
reinforcement to behave linear-elastically, the longitudinal
shear stress, , can be calculated using the expression:

= Vaddf Af (h x) / Ics ba

(Equation 12)

where
Vadd = difference between the ultimate shear force and the
applied shear force when the strengthening is installed
f
= short-term modular ratio of FRP to concrete
= Efd /Ec
Af
= area of FRP plate
x
= depth of neutral axis of strengthened section
= second moment of area of strengthened concrete
Ics
equivalent cracked section
ba
= width of adhesive layer
h
= total depth of the section
The longitudinal shear stress should be checked near to the
plate ends, where the shear force acting on the strengthened
portion of the member will be at its greatest. Equation 12
may be used in this position provided both the concrete in
compression and steel reinforcement are still behaving
approximately elastically. The longitudinal shear stress need
not, however, be checked within lt,max of the end of the plate,
where lt.max is determined in accordance with Equation 15.
Additionally, the longitudinal shear stress must be checked
where any changes in section properties occur, at positions
where there are discontinuities in shear force, such as at the
position of point loads, and at the location along the span
where the steel reinforcement stressstrain behaviour goes

35

Design guidance for strengthening concrete structures using fibre composite materials

from elastic to yielding. At this point the longitudinal shear


stress can increase dramatically because the steel reinforcement no longer contributes to the flexural stiffness of the
section and any increase in bending moment must be carried
by the FRP alone. The position in the span where the steel
first yields must be found, either by an iterative trial-anderror method or by calculating the strain profile along the
length of the beam. Once the position where the reinforcement yields is found, the longitudinal shear stress should be
found by calculating the stress ff in the FRP at this position
and at a position a short distance, y, further along the
section in the direction of increasing bending moment. The
difference in stress in the FRP, ff, along this length can then
be found and the resulting shear stress can be approximated
using the equation:

= tf

f f
y

(Equation 13)

This approach to calculation of longitudinal shear stress


should be used wherever the section properties are non-linear.
Anchorage design

lt,max = 0.7 (Efd tf /fctm)

(mm)

where
kb
= 1.06. [(2 bf /bw) / (1 + bf /400)] > 1.0

(Equation 16)
bf
= plate width (mm)
bw
= beam width or plate spacing for solid slab (mm)
tf
= plate thickness (mm)
Efd

= elastic modulus of the plate (N/mm2)

fctm

= tensile strength of concrete = 0.18 (fcu)2/3 (N/mm2)


(Equation 17)

(Ideally fctm should be obtained from pull-off tests on the


actual concrete.)
It is recommended that, where the FRP is curtailed in the
span, a minimum anchorage length of 500mm should be
provided. In situations where it is not possible to provide an
anchorage length in excess of lt,max, the bond force will be
less than Tk,max and may be calculated using the following
expression:
Tk = (Tk,max lt /lt,max) [2 lt /lt,max] (N)

In addition to maintaining low longitudinal shear stresses,


adequate FRP end anchorage must be provided. Work on end
anchorage lengths has been carried out by a number of
authors. The recommended approach is based upon the model
proposed by Neubauer and Rostasy(111).
Figure 23 illustrates the model. Also as described above, it
can be seen that the characteristic bond failure force, Tk,
increases with increasing anchorage length, lt, but that there
is a threshold anchorage length, lt,max, above which no increase
in the bond failure force is possible.

(Equation 15)

(Equation 18)

The anchorage design should be undertaken by determining


the point in the span where the FRP is no longer required.
From a sectional analysis, in accordance with the approach
described in Section 6.2, the force that will be developed in
the FRP if it is bonded to the member at this point should
then be determined. It is important to recognise that, although
the FRP may not be required at this point, the fact that it is
bonded to the concrete will nevertheless mean that some
force will be developed in it. The resulting FRP force should
be checked to ensure that it is less than the ultimate
anchorage capacity, Tk, and if so, an acceptable anchorage
design will result from extending the FRP by an anchorage
length beyond this point. If this condition is not satisfied
then the FRP should be extended further towards the support
or consideration given to using a thinner but wider FRP
laminate. Alternatively, consideration may be given to using
an anchorage device, provided that its capacity has been
proved by testing.

6.4 FLEXURAL STRENGTHENING WITH


NSM REINFORCEMENT
6.4.1 Introduction
Figure 23: Characteristic bond failure force vs anchorage
length.

The maximum ultimate bond force, Tk,max, and the corresponding maximum anchorage length, lt,max, needed to activate this bond force can be calculated using the following
expressions:
Tk,max = 0.5 kb bf (Efd tf fctm) (N)

36

(Equation 14)

Flexural strengthening can be achieved by bonding


pultruded strips or rods into slots cut in the surface of the
concrete. This application is termed Near-Surface-Mounted
(NSM) reinforcement, and has benefits where the exposed
concrete surface is to be trafficked or otherwise exposed to
potential damage. The technique is also applicable where the
surface of the concrete is undulating, or if there is excessive
laitance or a thin layer of poor quality concrete near the

Strengthening members in flexure

surface. The method also results in an increased bond area,


which helps to delay the onset of debonding type failure.
Installation is more costly than for externally bonded reinforcement, due to the need to cut the slots and prepare the
bond surface. Usually the technique would only be used where
externally bonded reinforcement is not a good technical
solution.

Most experiments to date have been performed on NSM bars


with circular or rectangular cross-section. Circular bar diameters range between 7 and 16mm while strips are usually
rectangular in shape of thickness less than 2mm. Most tests
on NSM bars have used carbon or glass FRP bars(117,118).
Tests have shown that higher average bond stresses are
obtained from bars made from carbon than those made from
glass(118) probably due to the higher stiffness of carbon.

6.4.2 Design basis


Other than for FRP curtailment (see Section 6.4.4), the basis
for design of NSM schemes is substantially the same as for
surface-mounted flexural strengthening. Appropriate allowance must be made for the fact that the FRP reinforcement is
located within the section, and will therefore be strained
slightly less than the surface of the concrete. Flexural design
methods and limits should therefore be as for surfacemounted reinforcement (see Section 6.2).
Material details and compatibility should be in accordance
with the manufacturers recommendations. Complete systems
should be adopted, since details, such as the preparation of
the surface after the groove has been cut, depend upon the
adhesive used. For example, clean dry surfaces are normally
required, but primers may be necessary before certain adhesives are applied to the concrete. The type and thickness of
epoxy to be used should be in accordance with the manufacturers recommendations. Since this may also be affected
by the location and orientation of the groove and specific
requirements of the structure (e.g. environmental exposure,
service conditions such as elevated temperatures) this should
be discussed with the manufacturer early in the design
process.
6.4.3 Bond behaviour

To improve the bond capacity, NSM bars have either a


prepared surface (by grit-blasting, abrasion or peel ply) or a
deformed surface, with ribs similar to deformed steel bars or
spiral surface deformations, as a result of the method of
manufacture. It has been observed that deformed bars perform
better in terms of bond than grit-blasted bars(118).
The size and shape of the groove into which the bar is placed
affect the mode of failure of the anchorage of an NSM bar. It
should be noted that other factors such as strength and
thickness of the adhesive surrounding the bar interact with
the size (depth and width) to determine the mode of failure.
Although cement-based adhesive materials can be used to
fill the grooves and surround the bar, experimentation has
shown that this gives lower average bond strengths than
epoxy adhesives, which have higher tensile strengths.
Expansive cement-based mortars should be avoided since
the expansion can introduce cracks and weaken the bond(119).
Furthermore, most of the existing experimental results on
NSM techniques have been obtained using epoxy adhesives.
Due to the experimental results noted above, and reflecting
the bars currently available in the UK market, guidance is
presented here for a subset of the possible NSM schemes.
6.4.4 Modes of failure

Curtailment and anchorage of NSM FRP differs from that of


surface-mounted FRP. The adhesive is a thicker block than
the thin layer used for surface-mounted strips. The interface
area between FRP and adhesive is potentially much smaller
than that between adhesive and concrete. Furthermore the
practicalities of surface preparation are likely to result in
different qualities of preparation on the sides of the slots than
on the bottom of the slot. For these reasons, some of the
debonding criteria used for surface-mounted plates are not
applicable to anchorage of NSM rods.

Both pull-out tests and beam tests have indicated that the
ultimate load carried by NSM bars increases with increasing
bond length. As for surface-mounted strips, there is a limit to
this length beyond which any further length increase does
not enhance the bond strength(117). However, the strength
arising from this limit is a much higher proportion of the
FRP capacity than is the case for surface-mounted strips, and
is correspondingly less likely to be a limiting design factor.

There are more variable factors for NSM applications than


for surface-mounted strips and plates. The key factors include
bar size and shape, bar material, bar surface preparation
(deformed or sandblasted), groove size, shape and surface
preparation, bonding agent (epoxy or cementitious material),
strength of concrete and location of existing bars. The effects
of these factors have been noted experimentally and are still
being actively researched. However, few experimental programmes have examined all these variables and how they
interact.

There are two main modes of failure associated with NSM


bars. Failure normally involves debonding, which occurs at
the adhesivebar interface or concreteadhesive interface,
although in some cases tensile failure of the FRP bars may
occur before bond failure. The two main failure modes are:

Adhesive splitting failure: splitting of the adhesive cover


surrounding the bar as a result of high tensile stresses that
are initiated at the FRP barepoxy adhesive interface
Concrete splitting failure: splitting of the concrete
surrounding the adhesive as a result of concrete at the
interface reaching its tensile strength. This mode normally
occurs when the adhesive strength is much higher than
the surrounding concrete.

37

Design guidance for strengthening concrete structures using fibre composite materials

These modes of failure are generally accompanied by pullout of the NSM bar along the interface where the failure is
initiated. Mixed mode failure involving a combination of the
two has also been reported(117120).
Adhesive splitting is predominantly governed by the thickness of the epoxy surrounding the NSM bar while concrete
splitting is governed by the groove size. By using adhesives
of high tensile strength, adhesive splitting failures, which
form with longitudinal cracking through the adhesive cover,
can be minimised(121). It is then possible to increase the
groove dimensions and minimise the induced tensile stresses
at the concreteadhesive interface thus preventing concrete
splitting failure.

published test data for the performance of NSM bar


anchorage with a wide range of properties. The guidance is
limited to the following:

6.4.5 NSM separation failure design


Shear-crack-induced separation, surface irregularity-induced
separation and maximum FRP strain should be checked as
for surface-mounted strips (Section 6.3.2). In checking irregularity, the profile of the as-installed bars should be considered, since NSM bars may be appropriate for an undulating
surface if this is corrected by the groove-cutting process.
Outside the anchorage length, the longitudinal shear stress
between the adhesive and concrete should also be checked as
for surface-mounted strips, except that in applying Equation
12 an equivalent value should be substituted for ba,
representing the useful perimeter of the groove. For a
rectangular groove this would normally be the minimum
width plus the minimum depth, i.e. only half of each side of
the groove is counted since the groove sides cannot normally
be prepared to as high a standard as an exposed face. If
special methods and particular care are used on the sides of
the groove, it may be appropriate to increase the useful
perimeter to the width plus twice the depth, i.e. the gross
perimeter of the groove. If there is a layer of weak laitance
near the surface of the concrete the value for the depth
should be reduced appropriately.
6.4.6 Anchorage design
Two main approaches have been developed in order to
predict the anchorage length required and the maximum
stress that can be carried by NSM reinforcement. In the first
approach data from bond slip tests are used to model the
bond behaviour of NSM anchorage(118). These equations can
predict the anchorage length required and the resulting stress
in the bar. In the second method an average bond stress is
assumed and the radial pressure exerted on the surrounding
concrete is related to the ultimate cracking stress in the
adhesive or concrete interface(117).
However, in view of the variability and relative novelty of
the technique, a cautious approach is suggested, which may
be relaxed if a more rigorous design method based on
specialist advice is carried out. The guidelines are presented
here for a subset of possible NSM schemes based upon

38

circular bars of up to 16mm diameter


epoxy adhesive used to bond bars into grooves
surface preparation of the bar (e.g. peel ply or a deformed
surface) such that manufacturers testing demonstrates
that the bar will not pull out from the adhesive used
grooves such that the installed bar is straight
grooves having square cross-section, with both width and
depth of groove at least twice the bar diameter
existing structural metallic reinforcement not intersecting
the groove
where grooves are spaced on the structure, the clear
spacing between grooves should be at least the width of
the groove and the last groove should be at least four
times the bar diameter from an edge of the structure
grooves having a surface preparation that provides a
rough gripping surface for bonding (i.e. not simply
diamond-sawn).

For conditions outside these criteria, specialist advice should


be sought. Detailed analysis as described in the references
may be required to confirm anchorage, or it may be
appropriate to undertake testing to confirm the assumed
performance.
To avoid adhesive splitting failure, bar force should be
limited to no more than:
Tnsm,ad = 0.1bbarperim lnsm f at for plain FRP bars (including
spirally wound and sand coated
bars)
(Equation 19)

Tnsm,ad = 0.3 bbarperim lnsm f at for deformed FRP bars


(Equation 20)

where
Tnsm,ad
bbarperim
lnsm
fat

= characteristic adhesive bond failure force


= perimeter of FRP bar
= anchorage length provided for NSM bar
= design adhesive tensile strength.

To avoid concrete splitting failure, the maximum ultimate


anchorage force, Tnsm,max, and corresponding maximum
anchorage length, lnsm,max, can be calculated from the following expressions:

Tnsm, max = 1.9 Efd Af bnotchperim f ctm

lnsm, max = 4.5

Efd Af
bnotchperim f ctm

(Equation 21)

(Equation 22)

Strengthening members in flexure

where
Tnsm,max = maximum NSM anchorage force (N)
lnsm,max = anchorage length required to generate Tnsm,max
Efd

= design FRP modulus of elasticity (N/mm2)


(mm2)

Af
= area of FRP
bnotchperim = effective perimeter of notch (mm) (making
allowance for surface preparation and/or weak
laitance layer)
= concrete tensile strength (N/mm2)
fctm
In situations where the maximum anchorage length is not
possible or necessary, the anchorage force generated by a
shorter length can be assessed from:
Tnsm = Tnsm, max

lnsm
l
2 nsm

lnsm, max
lnsm, max

6.6 THICK AND MULTI-LAYER


LAMINATES
The FRP strengthening systems available in the UK are generally pultruded plates of less than 2mm thick (1.2mm and
1.4mm are the most common) and fabrics with an effective
thickness of between 0.1mm and 0.3mm (see Appendix B).
Some manufacturers can supply thicker plates, either manufactured in a single process or by bonding together previously manufactured pultruded plates to produce a thicker
laminate. In addition, it is possible to bond plates in stacks
on site. Fabrics are laminated on to the structure by applying
successive layers of resin and fabric until the required
thickness (and hence strength) is obtained.

(Equation 23)

where
Tnsm = characteristic anchorage force for NSM
lnsm = anchorage length provided for NSM.
While the use of NSM strips, rather than rods, appears to be
an efficient strengthening technique, there is at present insufficient information available for specific design guidelines
for such situations to be included and specialist advice
should be sought.

6.5 FLEXURAL STRENGTHENING


PLATE LOCATION
Flexural strengthening is usually achieved by bonding plates
(or fabric) to the tensile face of the member being
strengthened (i.e. the soffit of a sagging simply supported
beam). This achieves maximum efficiency in use of material,
since the FRP is subjected to the maximum possible tensile
strain. However, in certain situations it may be appropriate
to bond plates to other parts of the section.
If this is undertaken, there is reduced utilisation of the FRP
as it is less highly strained since it is located nearer to the
neutral axis of the section. Other issues must also be
considered. Most significantly, if it is proposed to bond the
FRP to the sides of the beam, the strip will now be subjected
to bending about its strong axis, in which it may have
significant flexural stiffness. Although this may make a
small additional contribution to the flexural stiffness of the
section as a whole, the most significant effect is that
debonding of the plate may be precipitated, since in addition
to tensile peeling and longitudinal shear stress, the adhesive
interface is now also subject to transverse shear stresses.
There has been no significant research with plates in this
orientation. If this arrangement is considered, the adequacy
of the design should be confirmed by testing, unless the
beam is very deep and therefore the differential strain
between the edges of the plate relatively small

However, there are limitations to the thickness of FRP that


can be usefully employed. Frequently, the prevention of a
debonding failure of the FRP from the concrete will limit the
thickness, since stacking two laminates will almost double the
longitudinal shear stresses in the FRPconcrete adhesive bond.
Additional layers increase the number of potential failure
modes, since failure can occur in the adhesive between
layers and exacerbates potential failure within the FRP. If the
stacked layers are not of the same length (which is normal,
in order to reduce stress concentrations in the curtailment
zone), there are also additional anchorage zones that must be
checked for debonding.
It is preferable not to stack pultruded plates. In some situations (e.g. T-beams with narrow webs that have insufficient
width of soffit to accommodate the required areas of FRP) it
may be the only way that an otherwise suitable scheme can
be detailed. In these situations stacking of plates in situ, or
use of thicker laminated plates may be appropriate if the
following conditions are met:

The plates are intended by the manufacturer for such use,


and have sufficient inter-laminar strength that they will
not suffer internal shear failures.
The plates that are bonded to both sides have suitable
preparation to both faces; some plates incorporating peel
ply are manufactured with this on only one face, and
would not be suitable for stacking without further preparation work.
Peeling failure is checked at the curtailment of every
plate, and allowance is made in the design calculations
for the non-prismatic section.
The construction sequence is carefully specified to ensure
that bonding subsequent plates does not disturb or damage
the bond of the lower plates.

For stacks of plates formed in situ, the stack should not


normally be more than two high. There is some limited UK
experience with stacks three high, but in this case specialist

39

Design guidance for strengthening concrete structures using fibre composite materials

advice should be sought. For stacks formed by the manufacturer under factory conditions, higher stacks may be
possible. Testing may be necessary to demonstrate the performance of the stack, but it is unlikely that greater than 5mm
thickness will be useful, due to the limits of the bond
strength to the concrete.
When fabrics are used, multiple plies can be overlaid to
achieve the necessary strength of the FRP component. The
same overall limitations apply, but these limits will normally
be reached when many plies are installed. If large numbers
of plies are overlaid, it is likely to be the achievable quality
of workmanship that limits the design, and a trial installation
should be considered in order to demonstrate that a void-free
laminate can be produced under the site conditions relevant
to the particular project.

widths in steel-reinforced concrete structures is given in


Section Three of BS 8110: Part 2, and Section Five of BS
5400: Part 4. These methods can be adapted for FRPstrengthened structures simply by calculating the strain in
the tension steel, using a suitable concrete compression
stressstrain curve, and crack width under permanent load
and transient load separately. The FRP strengthening can
betaken into account by using the transformed area of the
FRP laminate in calculating the strain in the tension steel
under transient loading (unless the construction sequence is
such that the FRP laminate will be subject to permanent
loads, for example where surface finishes are applied over
the FRP laminate). The second moment of area of the section
should be determined assuming that the long-term modular
ratios of steel to concrete, e, and FRP to concrete, f, are
given by:

6.7 REDISTRIBUTION

e = Es /( Ec) or e = Es /((1/1+) Ec) (Equation 24)

Moment redistribution in any continuous concrete structure


relies on adequate rotation capacity of critical sections(122). If
a structure displays ductility, it will also display rotation
capacity. However, if a structure displays rotation capacity, it
will not automatically display ductility(123125). This is important in the context of FRP-strengthened concrete structures
because research has shown that ductility is usually limited
in such structures(126), even though adequate levels of rotation capacity often exist(127,128).

f = Efd /( Ec) or f = Efd /((1/1+) Ec) (Equation 25)

Therefore, moment redistribution into FRP-strengthened


zones is permitted up to a maximum of 30% if it can be
demonstrated that there exists sufficient rotation capacity
within the strengthened zone and within the surrounding
structure to allow such redistribution to be take place (see
Denton(129)).

where
=

creep coefficient.

It is worth noting that calculating crack widths is not as


straightforward as suggested here. This is because, when the
FRP strengthening system is placed on the surface, the crack
spacing is defined by the load on the unstrengthened
structure. The crack width due to live load would be significantly reduced due to the presence of the FRP strengthening. However, only a limited amount of experimental and
theoretical work defining the extent of this reduction in
crack width has been carried out. In the interim, therefore,
the procedure outlined above is recommended, which will
provide a conservative estimate.

6.8 SERVICEABILITY

Where a wet lay-up FRP strengthening system is used to


fully cover the concrete surface the durability of the concrete
structure may be significantly improved in the region of the
FRP strengthening. However, additional appropriate construction details, for example, to prevent the build-up of moisture
in the substrate adjacent to the wet lay-up FRP strengthening, should also be considered.

6.8.1 Crack widths

6.8.2 Deflections and material stresses

The causes of any existing cracking in the concrete substrate


should be ascertained, and resolved if possible, prior to
installation of the FRP strengthening system. Any such
cracks should be repaired prior to installation of the FRP
strengthening system, for example by resin injection; all
materials used should be compatible with the FRP
strengthening system.

For buildings, deflections due to the projected increased load


should not exceed the limits recommended in BS 8110. To
avoid excessive deformations in bridges, the stresses in the
steel reinforcement and concrete at working loads should not
normally exceed 0.8fy and 0.6fcu (or 0.6 times the worst
credible strength), respectively. See also Sections 5.6.5 and
5.6.6. Such deflections and material stresses can be
estimated using elastic principles. The presence of the FRP
strengthening will only reduce the live-load deflections
(assuming that the FRP is not prestressed and jacking is not
used). The material stress limits in Sections 6.8.3 and 6.8.4
are applicable. The equivalent transformed section for longterm loading should be obtained by assuming that the

However, moment redistribution out of FRP-strengthened


zones relies on a substantially higher level of rotation capacity being available in the strengthened zone. Therefore,
moment redistribution out of FRP-strengthened zones is not
recommended.

Normally, crack widths will not be excessive providing the


FRP strengthening system has been properly installed.
Where necessary crack widths at service loads should be
checked against the limits recommended in BS 8110 or BS
5400: Part 4, as appropriate. Guidance on calculating crack

40

Strengthening members in flexure

modular ratios of steel to concrete, e, and FRP to concrete,


f, are given by Equations 24 and 25, respectively, and for
short-term loading are based on the short-term elastic moduli
of the concrete and steel.

It is recommended that the maximum stress in the FRP at


service loads, as a proportion of the design strength, should
not exceed the values given in Table 9.
Table 9: Maximum stress under service loads to avoid stress rupture as a proportion of design strength (%).

6.8.3 Fatigue
For bridges, the designer should consider the effect of
repeated live loading on the fatigue strength of the steel
reinforcement and the FRP. Checks for fatigue failure should
be carried out in accordance with the recommendations in
Clause 4.7 of BS 5400: Part 4. The stress range in the FRP
should be limited to the appropriate value given in Table 8.
The failure mode for fatigue is typically yield of the steel
followed by spalling at the concrete soffit and debonding of
the FRP.
Highways Agency Advice Note BA 30(9) controls the fatigue
behaviour of steel plate bonding applications by limiting the
cyclic stresses that may be applied to the steel plate.
Table 8: Maximum stress ranges as a proportion of the design
ultimate strength (%).

Material

Stress range (%)

Carbon FRP

80

Aramid FRP

70

Glass FRP

30

Lap splices of FRP plates


Wherever possible, lap splices of FRP plates should be
avoided in regions where the fatigue stress range is high. In
addition, lap splices in adjacent plates should be staggered
and alternated to avoid the concentration of lap splices in a
single region.
Preliminary research(130) has shown that the fatigue life of a
lap splice is dependent on the length of the lap, among other
factors. Therefore, the minimum length of any lap splice
should be 300mm or the manufacturers minimum recommended lap splice length, whichever is the greater.
Delaminations/voids
Concrete Society Technical Report 57(5) provides guidance
on the acceptable limits of delamination and voids in an
adhesive bond.
6.8.4 Stress rupture
Rupture of the FRP may occur at service loads due to sustained stresses in the material. Applications where sustained
stresses may be present in the FRP include:

temporary dead load removal by jacking, followed by


FRP strengthening
physical removal of dead load, followed by FRP
strengthening and then reinstatement of dead load.

Material
Carbon FRP
Aramid FRP
Glass FRP

Maximum stress (%)


65
40
45

6.8.5 Strengthening under non-static live load


One of the great benefits of using FRP for retrofit strengthening is the ease, speed and short period of time required for
the works. In order to maximise these benefits during the
strengthening of concrete bridges, it is clearly desirable that
traffic be allowed to flow over the bridge during the works.
In the USA, strengthening under such live load is permitted,
as long as vehicle speeds are controlled(131).
The obvious caveat to strengthening under live load is the
effect that intermittent loading has on the curing process in
the adhesive. Hejll et al.(132) have carried out laboratory tests
on concrete bridge girders strengthened under static conditions and under simulated live-load conditions, in which a
load causing 60% of the yield strain in the steel reinforcement was applied cyclically every 108 seconds. Their tests
showed that there was effectively no difference in ultimate
capacity and behaviour between bonded systems (either
laminate or NSM) that had been installed under static or under
simulated live-load conditions. What is more, the level of
strengthening was 80150% of the original capacity of the
unstrengthened bridge girders, adding confidence to their
claim that traffic need not be stopped during FRP strengthening works on bridges.
This tends to confirm work by Barnes and Mays(133) who
conducted tests on strengthened beams with a 1Hz cyclic
load during curing of between 20 and 170 microstrain. Their
research showed that for strengthening concrete beams under
live load there was no effect on the ultimate strength. This was
because failure in their test specimens was due to the concrete cover pulling off, rather than failure of the adhesive.
However, further tests designed to investigate failure in the
adhesive itself showed a significant reduction in strength if
the adhesive interface governed.
It is therefore recommended that the strength of the adhesive
be reduced in accordance with Table 10. If this reduced
adhesive strength is less than the design strength of the
concrete, then the live load during adhesive curing should be
restricted. This should be taken into account when calculating achievable anchorage force in NSM-strengthened
structures.

41

Design guidance for strengthening concrete structures using fibre composite materials

Table 10: Reduction in strength of adhesive for given live-load


strains at FRPconcrete interface (%).

Live-load strains at
FRPconcrete interface
during curing (106)
20
50
100
150
200

Reduction in strength of
adhesive (%)

10
12
16
22

32

6.9 STRENGTHENING PRESTRESSED


STRUCTURES
Many of the same principles outlined in the previous
Sections of this Chapter can be applied to the strengthening
of prestressed concrete structures. However, note that service-

42

ability issues govern prestressed design whereas these guidelines focus predominantly upon ultimate limit state. There are,
therefore, a number of factors to consider when strengthening a prestressed structure such as:

the need to accurately assess current and future stress


states
the sensitivity of the strengthened section to the initial
stress state compared with that of an uncracked reinforced concrete section
the lower ductility of prestressing tendons compared with
reinforcing steel
different anchorage behaviour due to reduced cracking
preventing the full generation of anchorage force.

A fuller discussion of strengthening prestressed concrete


structures is given in the FIB Technical Report(90).

Strengthening members in flexure

43

Design guidance for strengthening concrete structures using fibre composite materials

44

Strengthening members in flexure

45

Design guidance for strengthening concrete structures using fibre composite materials

46

SHEAR STRENGTHENING

7.1 INTRODUCTION
Externally bonded FRP laminates and fabrics can be used to
increase the shear strength of reinforced concrete beams and
columns. FRP may be bonded to the concrete in various
configurations. Ideally FRP should be wrapped around the
whole perimeter of the member (fully wrapped). Alternatively, it can be applied only to the sides of the member
(side-only) or to the sides and the tension face of the member
(U-wrapped). Figure 24 shows examples of possible FRP
shear strengthening configurations. This Chapter focuses on
rectangular beams and columns. Guidance on strengthening
of circular columns in shear is included in Section 8.3.4.

can therefore be governed by separation of the FRP from the


concrete, and it is not sufficient to assume that fracture of the
FRP will occur. Such separation is typically associated with
the propagation of a failure plane in the concrete close to the
surface.
Side-only or U-wrapped members will be more prone to
separation failures than fully wrapped members. Full wrapping is therefore preferable and should always be used when
it is feasible. However, it is generally not practicable for
beams, because the top of the beam is inaccessible. In most
cases it will be possible to fully wrap columns.
The behaviour of reinforced concrete in shear is complex.
Furthermore, considerably less research has been undertaken
into FRP shear strengthening than flexural strengthening. It
is therefore appropriate at present to adopt a cautious design
procedure for shear strengthening. The design procedure given
is based upon that proposed by Denton et al.(112). In developing their proposals, they reviewed numerous alternative
design approaches and provide a detailed justification for
their proposed method.

Side Only

U-Wrapped

Fully Wrapped

(n=2)

(n=1)

(n=0)

Figure 24: Shear reinforcement configurations.

The orientation of the FRP fibres can affect the performance


of the strengthening system. Theoretically fibres that are
inclined to resist the formation of shear cracks can be more
effective than fibres aligned perpendicular to the longitudinal axis of the member. However, if the shear force
direction can reverse, or if the FRP is partially or fully
wrapped around the beam, systems with fibres aligned
perpendicular to the longitudinal axis of the member are
more convenient and are typically used in practice.

The majority of experimental testing of reinforced concrete


members strengthened in shear has used carbon rather than
aramid or glass fibre. Although the underlying principles
should be common to all materials, the design procedure
presented below is best suited to designs using carbon FRP.
The approach should be conservative if applied to aramid or
glass FRP, and in some cases may be significantly so.
As with flexural strengthening, the assumptions made in the
design should be reflected in the installation work on site. It
is therefore important to consider the issues outlined in
Section 6.1 in developing the design for shear.

7.2 DESIGN PROCEDURE


7.2.1 Maximum shear capacity

In understanding the behaviour of FRP strengthening in


shear, it is important to recognise that the bond behaviour of
FRP differs markedly from conventional embedded steel
reinforcement. As discussed in Section 6.3, it has been found
in tests on the anchorage of externally bonded FRP that
beyond a limiting bonded length, no further increase in the
ultimate anchorage load-capacity occurs with increasing
bonded length. This maximum anchorage capacity can be
very much less than the ultimate tensile capacity of the FRP.
The contribution that the FRP makes to the shear capacity

Irrespective of the amount of conventional or FRP shear


reinforcement provided, the shear strength of a section can
be limited by a diagonal compression shear failure in the
concrete. Such a failure can be avoided by limiting the maximum shear stress in the concrete. According to Clause 3.4.5.2
of BS 8110, the maximum permissible shear stress, vmax,
should be taken as the lesser of 0.8fcu or 5N/mm2, whatever
shear reinforcement is provided. Similar limits are included
in BS 5400 Part 4, where vmax is the lesser of 0.75fcu or

47

Design guidance for strengthening concrete structures using fibre composite materials

4.75N/mm2. It is recommended therefore, that such a criterion


be used in the design of FRP shear-strengthened members.
The maximum allowable design shear force due to ultimate
loads, VR,max, at any cross-section, is thus obtained from:
VR,max = vmax b d

(Equation 26)

where
b
= width of section
d
= effective depth of section
vmax = maximum concrete shear stress
(determined from the appropriate Standard)
Tests on FRP-strengthened beam and slab structures have
indicated that the shear capacity of such structures can be
limited by the longitudinal shear capacity at the interface
between the beam and slab. This interface should therefore
be checked to ensure its adequacy, using conventional design
or assessment Standards.
7.2.2 FRP shear strengthening design
The ultimate shear capacity of an FRP strengthened beam
can be expressed as:
Vu = Vc + Vs + Vf

The notation is illustrated in Figure 25.

(Equation 27)

where
Vc = contribution from the concrete to the shear capacity
(N)
Vf = contribution from the FRP to the shear capacity (N)
Vs = contribution from the steel to the shear capacity (N)
Vu = ultimate shear capacity of FRP strengthened section
(N).
Vc and Vs can be determined from conventional design standards, such as Clauses 3.4.5.3 and 3.4.5.4 of BS 8110: Part 1:
1997. For structures designed to the 1985 edition of BS
8110, it would be more appropriate to assume a steel stress
of 0.87fy, rather than 0.95fy, as in the 1997 edition. Alternatively, equivalent criteria are included in BS 5400 Part 4.
Assuming a 45 shear crack, it follows from equilibrium that
the contribution of the FRP to the shear capacity is given by:
n

d f lt ,max
3

(cos + sin ) (Equation 28)


Vf = Efd fse Afs
sf

where
n
= 0 for a fully wrapped beam, 1.0 when FRP is bonded
continuously to the sides and bottom of a beam and
2.0 when it is bonded to only the sides of a beam

= angle between the principal fibres of the FRP and a


line perpendicular to the longitudinal axis of the
member. is positive when the principle fibres of
the FRP are rotated away from the direction in
which a shear crack will form

48

fse = effective strain in the FRP for shear strengthening


Afs = area of FRP (mm2) for shear strengthening measured perpendicular to the direction of the fibres.
When FRP laminates are applied symmetrically on
both sides of a beam, Afs is the sum of the areas of
both laminates, i.e. Afs = 2 bf tf
bf
= width of the FRP laminate (mm) measured perpendicular to the direction of the fibres. For continuous
FRP sheet, sf is taken as 1.0 and bf is taken as cos
df
= effective depth of the FRP strengthening, measured
from the top of the FRP to the tension reinforcement
(mm)
Efd = design tensile modulus of the FRP laminate
(N/mm2) (see Section 5.6.3)
lt,max = anchorage length required to develop full anchorage
capacity (see Section 6.3) (mm)
sf
= longitudinal spacing of the FRP laminates used for
shear strengthening (mm). For continuous FRP
sheet, sf is taken as 1.0 and bf is taken as cos
tf
= thickness of the FRP laminate (mm).

Afs = 2bf tf

h
sf

bf

FRP Laminates on both sides

Figure 25: General notation for shear strengthening.

The effective strain in the FRP, fse, accounts for the variation in strain in the FRP along the shear crack when the ultimate limit state is reached. It should be taken as the minimum
of:
(i) fd / 2
(ii) 0.64

f ctm
Efdt f

(iii) 0.004
where
fctm = tensile strength of the concrete (N/mm2)
fd = design ultimate strain capacity of FRP (see Section
5.6.4)
The first strain limit of half the ultimate strain capacity
represents the average FRP strain when fracture of the FRP
occurs. Alternative limits have been suggested for this condition. Chen and Teng(134) propose half the ultimate strain capacity, while Tljsten(92) proposes 0.6 times the ultimate strain

Shear strengthening

The second strain limit corresponds to debonding of the FRP,


and is based on Neubauer and Rostasys anchorage model(111),
as described in Section 6.3. In other design approaches, FRP
separation has been considered primarily for FRP bonded
either to the sides of beams or to the sides and the tension
face of beams. Here it is recognised that this condition should
also be applied to fully wrapped beams to ensure that the
integrity of the concrete is maintained and Vc does not diminish before Vf reaches its design value, as explained by
Denton et al.(112). For small beams such an approach may be
conservative, but importantly it should be safe for the cases
most frequently encountered in practice.

Figure 27 shows a comparison of experimental measured


values of Vf and those determined using the design method
described above (after Denton et al.(112)), with all partial
safety factors taken as unity. There appears to be reasonable
agreement. Furthermore, the substantial majority of cases lie
on the safe side, with the experimentally measured values
exceeding those determined using the proposed design
method.

Vf (Design Method)

capacity. It appears that Tljstens limit applies when the


behaviour of the member is predominantly elastic and that
Chen and Tengs limit is more suitable when the behaviour
of a member is characterised by rigid body movements of the
regions of the member either side of a shear crack. The lower
of the two values has been adopted.

200
180
160
140
120
100
80
60
40
20
0

Side and U-Wrapped


Fully Wrapped
Design = Exp

The final 0.004 strain limit was proposed in early design


methods to ensure that the concrete integrity is maintained.
This convenient rule of thumb appears to have limited
rational justification, and as is shown by Denton et al.(112),
does not necessarily prevent the development of wide cracks.
It is retained because it seems sensibly cautious to do so.
In Equation 28, the effective depth is reduced by a length
equal to (n/3)lt,max. This adjustment accounts for the reduction in force that can be sustained by the FRP in the anchorage regions. The Neubauer and Rostasy anchorage model, as
described in Section 6.3, assumes a parabolic variation of
stress with distance (see Figure 26). The force corresponding
to the area under the stress curve in the anchorage region is
therefore only 2/3 of the maximum stress multiplied by the
anchorage length. This reduction in the FRP contribution can
be modelled by subtracting (n/3)lt,max from the effective
depth, as in Equation 28. The adjustment is made at the top
for U-wrapped beams (n=1) and at the top and bottom for
beams with FRP bonded only to the sides (n=2). No adjustment is necessary for fully wrapped beams (n=0).
E fd = 230GPa
fd = 0.015
f ctm = 3MPa

0.25
0.2

t f = 0.5mm

Parabolic Curve

0.15

t f = 1mm

100

200

300

400

500

600

V f (Experimental)

Figure 27: Experimental verification of design method (after


Denton et al.(112)).

7.3 SPACING OF FRP STRIPS


As in the case of steel shear reinforcement, the centre-tocentre spacing of strips of FRP should not be so wide as to
allow the full formation of a diagonal crack without intercepting a strip. In addition, Equation 28 is based on the
approximation that the FRP contribution to the shear
resistance is distributed across the whole crack, rather than
in discrete locations, which becomes invalid at large strip
spacings. For these reasons, if strips are used, their centre-tocentre should not exceed the least of:
(i)
(ii)
(iii)

0.8df
df (n/3)lt,max
bf + df /4

where the variables are as defined after Equation 28.


Alternatively, the contribution of FRP strips to shear capacity may be evaluated with a rigorous analysis, accounting
for the critical location for a shear crack and the effect of
anchorage of the FRP strips. If this approach is used, the
limits on strip spacing in (i) to (iii) need not apply.

0.1

t f = 1.5mm
0.05
0
0

100

200

300

400

500

600

Bonded Length (mm)

Figure 26: Typical variation in ultimate strain capacity with


bonded length (after Neubauer and Rostasy(111)).

7.4 ADDITIONAL AXIAL FRP


Using the truss analogy it can be shown that beam and
column elements subjected to a shear force will experience
axial tensile forces, additional to those due to bending.
Additional axial reinforcement may therefore be required
when strengthening for shear.

49

Design guidance for strengthening concrete structures using fibre composite materials

The standard method is to simply extend the axial FRP


reinforcement a distance of half the effective depth beyond
the point at which it is no longer required for bending.
If there is insufficient FRP to sustain the additional axial
tensile forces then extra axial FRP, Afa, should be provided.
The amount of additional FRP required can be determined
from:
Afa = Vs / 2ff

(Equation 29)

where
Vs = shear force due to ultimate loads
= stress in the FRP at the same location determined
ff
from a flexural analysis.

50

Clearly these approaches are not relevant when no axial FRP


is present for bending. In this case, the ultimate bending
capacity of the member should be re-evaluated assuming the
area of each axial reinforcing bar between the tension face
and the mid-depth of the section is reduced by an amount
equal to:
Vs
2ne f y ms

(Equation 30)

where
ne = total number of effective axial reinforcing bars
between the tension face and the mid-depth of the
section.
Any shortfall in bending capacity should be compensated for
by providing axial FRP reinforcement.

Shear strengthening

51

Design guidance for strengthening concrete structures using fibre composite materials

52

STRENGTHENING AXIALLY
LOADED MEMBERS

8.1 INTRODUCTION
Concrete columns in existing structures such as bridges and
buildings may require upgrading to enhance the following
properties:

axial load capacity


flexural capacity
shear capacity
ductility.

Increased axial load capacity, for example, may be needed


for compression required to carry higher loads than originally envisaged or where the loading requirements have
changed. Enhanced flexural strength may be required for
bridge supports that:

are not capable of fully sustaining design loads from


heavy vehicle impact
have insufficient lap lengths
have incorrect termination of longitudinal reinforcement.

Some columns designed to older codes may be incapable of


withstanding the large horizontal displacement that occurs
between member ends during an earthquake. They may therefore require ductility enhancement in order to hold the cover
concrete in place and prevent buckling of longitudinal reinforcement under axial load. Shear strength must also be
considered in any proposed column upgrading.
Where a deficiency exists, upgrading can be achieved by
bonding layers of axial and/or hoop FRP to the column
perimeter.
Generally, bonding axial FRP to the column surface enhances
the flexural strength of the member. Hoop wrapping will also
be necessary and should be placed over the axial FRP. The
hoop wrapping increases the concrete compressive strain
capacity, which can significantly improve the efficiency of
the strengthening design. It also prevents buckling of the
axial fibres, potentially enabling them to contribute in
compression. However, this contribution will be small and is
therefore generally neglected.
Bonding hoop FRP to the column surface enhances axial
load capacity and ductility of columns. The hoop FRP resists
lateral deformations due to the axial loading, resulting in a
confining stress to the concrete core, delaying rupture of the
concrete and thereby enhancing both the ultimate compressive strength and the ultimate compressive strain of the con-

crete. As noted in Section 4.2.2, this process is significantly


more efficient with circular than with square or rectangular
columns. This is because, with the latter, the confining action
is mostly concentrated at the corners. Measures for, and limitations of, strengthening columns of non-circular cross-section
are discussed in Section 8.6. It should be noted that for the
case of concentrically loaded columns, the bond is not
critical since the composite does not need to transfer forces
along the FRPconcrete interface. However, bond becomes
important when moment is also applied to the column and
anchorage of the localised confining hoop FRP is required.
Column strengthening is normally carried out using fabric,
which may be applied dry or be preimpregnated with an
epoxy resin. The use of preformed shells made from a range
of fibre types, including glass and carbon, is another option.
However, at present, most of the studies that have been
carried out using FRP shells have concentrated on their
potential for new construction, where the FRP acts both as
permanent formwork to the wet concrete and as external
reinforcement, rather than in repair or strengthening work.
The following sections deal with the design of circular
columns for enhanced compressive strength, flexural strength,
shear strength and ductility. A feature common to the design
procedures discussed is the stressstrain model for FRPconfined concrete, which is examined next.
It should be noted that the basic principles of strengthening
columns given in this Chapter are applicable also to
strengthening with shells. However, this is a more complex
design process that includes aspects such as the performance
of the grout annulus and is beyond the scope of this Report.

8.2 STRESSSTRAIN MODEL FOR FRPCONFINED CONCRETE


Concrete in circular columns confined by hoop FRP displays
an approximately bilinear stressstrain response, as shown in
Figure 28. Initially, the behaviour is similar to that of plain
concrete since the FRP exerts a limited confining pressure on
the concrete. As the axial stress increases, however, the rate
of lateral deformation of the concrete also increases, which
results in a concomitant reduction in stiffness of the concrete. Once the concrete reaches the ultimate strain limit of
unconfined concrete, typically 0.0035, the material becomes
highly fissured and the confinement provided by the FRP is
fully activated. At this stage, the stressstrain response
becomes approximately linear with a slope dependent upon
the stiffness of the hoop FRP.

53

Design guidance for strengthening concrete structures using fibre composite materials

A large number of researchers have developed theoretical


stressstrain models validated by experimental testing(136139).
While these models are reported to show quite good agreement with test results(140), they are in some instances rather
complex and can be unconservative. There is also a need to
accurately predict ultimate stress and strain conditions. A
recent model developed by Lam and Teng(141) captures the
main aspects of the behaviour of FRP-confined circular
concrete columns in a simple form. The model has been
calibrated against all the currently available test data.

Figure 28: Idealised stressstrain curve for FRP-confined


concrete.

Several authors have proposed models that attempt to predict


the compressive stressstrain behaviour of confined concrete.
A large body of this work actually relates to steel rather than
FRP-confined concrete, with most of the formulations being
founded on the pioneering work of Richart et al.(135) on
triaxially confined concrete. Based on tests on cylindrical
specimens subjected to constant hydrostatic pressure, these
authors showed that both axial strength and ductility of the
concrete increases with increasing confinement pressure.
Further analysis of their data revealed that the compressive
strength increase could be predicted using the expression:
fccd = fc0 + 4.1 fr
where
fccd =
fc0 =
=
fr =

(Equation 31)

fcc = Eccc (Ec E2)2cc2/4fc0 for 0 <


cc < t
(Equation 32)
and
fcc = fc0 + E2 cc

confined concrete compressive strength


unconfined concrete compressive strength
0.67 fcu / mc
confinement pressure

Later work on the behaviour of steel-confined concrete


tended to confirm the validity of Equation 31. This is somewhat surprising considering the differences between the two
modes of confinement actually used. Richarts test specimens were subjected to active confinement due to the hydrostatic pressure, which remained constant throughout the test.
However, lateral steel reinforcement provides passive confinement since the confining stresses only develop as a result
of the lateral expansion of the concrete. In this case, the
confining stress is neither uniform nor constant.
Over recent years, certain disadvantages of steel confinement have emerged, such as corrosion, incompatibilities
between the modulus of elasticity and Poissons ratio of steel
and concrete, and aesthetics, which have led researchers to
consider FRP composites as an alternative confinement
medium. Strength models for FRP-confined columns, similar
to Equation 31, have been proposed by a number of
researchers. While these models may be adequate for predicting ultimate strength, they grossly underestimate ultimate
strain. Consequently, there are large discrepancies between
the actual and predicted stressstrain responses.

54

The model is made up of two parts, an initially parabolic


section, similar to that of unconfined concrete, followed by
a straight line section, the slope of which is dependent upon
the level of confinement. The initial slope of the parabolic
section is the same as that for unconfined concrete. The
parabola and straight line meet at the same slope and the
projection of the linear portion intercepts the stress axis at
the unconfined strength, fc0, as shown in Figure 29. The
model converges to a stressstrain model similar to that in
BS 8110 for the unconfined case. The model is defined as
follows:

for t <
cc < ccu
(Equation 33)

where
t
= position of transition region between parabola and
straight line
= 2fc0 /(Ec E2)
(Equation 34)
E2 = slope of linear portion of confined stressstrain
curve
= (fccd fc0)/ ccu
(Equation 35)
Ec = initial modulus of elasticity of concrete

cc
fcc
ccu
fccd

5.5(fcu/mc) (kN/mm2)
confined concrete axial strain
confined concrete axial compressive stress
confined concrete ultimate axial strain, given by
Equation 40
= confined concrete ultimate strength, given by
Equation 39.
=
=
=
=

It is clearly evident that in order use this model, the ultimate


design failure stress, fccd, and ultimate compressive failure
strain, ccu, of the concrete must be defined. Suitable models
are given in Section 8.3.2.
It should be noted that the proposed model is only suitable if
the strength of the confined concrete increases with
increasing strain. For low levels of confinement, following
the initial parabolic increase, the stress may decrease with a

Strengthening axially loaded members

further increase in strain and, therefore, the ultimate strength


may be lower than the peak strength. It is therefore recommended that this stressstrain model only be used when the
following condition, based upon the work of Xiao and
Wu(142), is met:

Compliance with relevant serviceability limit states, such


as axial shortening, lateral deformation, loss of strengthening effectiveness, fatigue and creep rupture, should be
investigated.

8.3.2 Tensile rupture of FRP

2tf Efd
> 0.183
Df c02

(mm2/N)

(Equation 36)

where
tf =

thickness of FRP (mm)

Efd =
D =

design modulus of elasticity of FRP (N/mm2)


diameter of column (mm)

If the condition is not met then no strength enhancement


should be assumed and the stressstrain model in BS 8110 or
BS 5400 for unconfined concrete should be assumed.
fcc

Generally it has been assumed that compression members


strengthened by hoop wrapping will fail if the circumferential stress in the composite exceeds its design tensile
stress capacity. On this basis, various equations have been
proposed for predicting the strength of FRP-confined
columns. A number of these equations are actually based on
the failure stress criterion for hydrostatic pressure proposed
by Richart et al.(135), given in Equation 31, and take the
following form:
fccd = fc0 + k fr

(Equation 37)

where the confining pressure, fr, is given by:

fccd
Confined concrete

fr = 2ffd tf / D

E2

(Equation 38)

1
fc0

and
k = confinement effectiveness factor.

Unconfined concrete

Ec
t

0.0035

ccu

cc

Figure 29: Stressstrain model.

It should also be noted that Equations 32 and 33 are both


calibrated against test results obtained by subjecting
cylindrical specimens to substantially concentric compression. In most practical situations, columns are subject to both
axial load and flexure, either due to load eccentricity or
applied moments. In particular, columns in bridge structures
can be loaded horizontally as a result of vehicle impact. In
this case, the column must sustain a combination of axial
load and bending moment. Methods for analysing sections
under combined axial and flexural loads are detailed in
Section 8.4.

8.3 COMPRESSION
8.3.1 Introduction
As previously mentioned, providing hoop FRP around the
perimeter of the column can increase the compressive strength
of circular columns. All members strengthened in compression should meet the following conditions:

Tensile rupture of the FRP should be considered.


Failure of the FRP jacket at lap joints should be
examined.
The shear capacity of the column should be checked.

Various values and expressions have been suggested by


researchers for the value of the effectiveness factor k
(typically less than the value of 4.1 given in Equation 31),
based on experimental results(137,139).
While such equations are reported to give good correlation
with experimental results, they do not satisfy lateral strain
compatibility requirements and are therefore deemed unsuitable. According to Lillistone and Jolly(143) this problem
would be overcome if the failure criterion were based on the
confining stiffness rather than the confinement pressure.
Such an approach would offer the following advantages:
The failure criterion does not require prior knowledge of the
lateral expansion of the concrete core. Unlike tensile strength,
the elastic modulus of fibres is not reduced by mechanical
abrasion during manufacturing processes.
On this basis, they recommended that the design strength of
concrete-filled filament-wound glass FRP circular tubes, fccd,
should be estimated using:
fccd = fc0 + 0.05(2tf/D) Efd

(Equation 39)

Comparative studies by Lillistone(140) show good agreement


with the results presented by Samaan et al.(137) for concretefilled E-glass-fibre filament-wound tubes, Howie and
Karbhari(136) and Picher et al.(138) for carbon-fibre-wrapped
concrete cylinders, and Saafi et al.(139) for concrete-filled Eglass and carbon-fibre filament-wound tubes.

55

Design guidance for strengthening concrete structures using fibre composite materials

In view of the above, it is recommended that Equation 39 be


used to calculate the axial failure stress of FRP-confined
concrete. It is worth noting that this equation is also based on
the characteristic unconfined cube strength of concrete and
that a partial safety factor for concrete of 1.5 is assumed.
The corresponding strain at rupture is required for developing the stressstrain curve in Section 8.2. Studies have
shown that the behaviour of confined concrete depends not
only on the confining pressure but also the stiffness of the
confining FRP(137). As a result, ultimate strains are different
for different confining FRP materials, even when the confining pressure is the same. Lam and Teng(141) have shown
that correlation with experimental results is poor if stiffness
of the confining FRP is neglected in the model of ultimate
strain. Based on experimental results, Lam and Teng have
developed an empirical equation for ultimate concrete strain
that takes into account the stiffness of the FRP as follows:

2E t
ccu
= 1.75 + 12 fd f
c0
E0 D

1.45

0.6 fd

c0

(Equation 40)

where
E0 = secant modulus of concrete = 0.67fcu/(mc c0)
c0 = axial strain in unconfined concrete at peak stress =
2.4104(fcu/mc)
fd = design ultimate strain of FRP.
It should be noted, however, that at concrete compressive
strains of over approximately 0.01, the concrete will have
been crushed and have lost all cohesion. It is therefore
recommended that if the ultimate strain, ccu, is greater than
0.01, then the failure stress should be taken as the value for
fccd corresponding to the value of cc=0.01 from the
stressstrain curve, rather than the failure stress at rupture of
the FRP.
8.3.3 Lap joint failure
Failure of the FRP jacket can occur at lap joints due to
debonding, if the lap length is inadequate. This type of
failure is brittle and can be avoided simply by providing an
adequate lap length. The actual length of overlap required is
likely to vary between strengthening systems and so it is
recommended that individual manufacturers are consulted.
Where necessary, independent testing should be carried out.
When two or more plies of FRP are applied to a column, the
lap joints should be arranged so they are staggered evenly
around the column perimeter, as shown in Figure 30. The
minimum overlap for fabric materials, in the direction of the
fibres, should be in accordance with the manufacturers
recommendations, but not less than 200mm.

column

column

Figure 30: Laps in columns.

8.3.4 Shear
The presence of hoop FRP can increase the shear strength of
concrete columns. Guidance on shear-strengthening of
square or rectangular beams and columns is included in
Chapter 7. The guidelines included in this Section relate to
circular columns.
The maximum shear strength of the section should be
determined in accordance with Section 7.2.1. This capacity
gives the upper limit on the degree of strengthening that can
be achieved.
The ultimate shear capacity of an FRP strengthened column
can be expressed as:
Vu = Vc + Vs + Vf
where
Vc =
Vf =
Vs =
Vu =

(Equation 41)

contribution from the concrete to the shear capacity


contribution from the FRP to the shear capacity
contribution from the steel to the shear capacity
ultimate shear capacity of FRP strengthened section.

As described in Section 7.2.2, Vc and Vs can be determined


from conventional design standards with account taken of
the material partial safety factors appropriate to the original
date of construction. Additional guidance on evaluating the
capacity of circular columns is given by Clarke and
Birjandi(144).
As discussed in Section 7, for rectangular or square sections
it is important to take account of debonding of the FRP in the
design of shear strengthening, even if the member is fully
wrapped. However, for circular members, the significance of
debonding is reduced because the development of tensile
stresses in the hoop FRP tends to improve the bond behaviour
by providing a lateral confining pressure.
For continuous hoop FRP wrapped around a circular column
with fibres aligned perpendicular to the longitudinal axis of
the member, Vf is given by:
Vf = (/2) tf d Efd fse

56

Lap to be
determined

(Equation 42)

Strengthening axially loaded members

where the effective strain in the FRP, fse, should be taken as


the lesser of:
(i)
or
(ii)

fd / 2
0.004

and
fd = design ultimate strain capacity of FRP
d
= effective depth (distance from the extreme compression fibre to the centroid of the tension reinforcement)
Efd = design tensile modulus of the FRP
tf
= thickness of the FRP.
As discussed in Section 7.4, additional axial reinforcement
may be required when strengthening for shear. Section 7.4
also outlines how this area of reinforcement can be determined.
8.3.5 Serviceability
Axial shortening/lateral deformation and loss of strengthening effectiveness
Axial shortening due to projected load increases will give
rise to lateral deformation of compression members. This
deformation, if excessive, may cause problems of appearance, damage to brittle finishes and/or loss of structural
efficiency. Also, at service loads the maximum compressive
strain in the concrete should not be excessive otherwise loss
of confining pressure due to accidental damage, fire,
vandalism, etc., may result in brittle collapse, because the
concrete is fissured. To prevent the possibility of either
problem arising, it is recommended that the axial compressive strain of the concrete should not exceed 0.0035 under
working loads.
Fatigue
For bridges, the designer should consider the effect of
repeated live loading on the fatigue strength of the FRP.
Checks for fatigue should be carried out in accordance with
the recommendations in BS 5400: Part 4. The stress range in
the FRP should be limited to the appropriate values given in
Table 8 of this Report.
Stress rupture
Rupture of the FRP may occur at service loads due to the
sustained stresses that exist in the material. This type of
failure can be avoided simply by limiting the stress level in
the FRP. It is therefore recommended that the stress in the
FRP should not exceed the values given in Table 9 of this
Report.

8.4 FLEXURE
8.4.1 Introduction
Bonding axial FRP over-wrapped with hoop FRP to column
surfaces can enhance the flexural strength of columns. The
main benefit of the axial FRP is to increase the flexural
strength of the member, and the problem in design is to determine the thickness of axial FRP fibre required to resist the
combined design axial load and moment. The hoop wrapping confines the concrete, increasing its compressive
strength and strain to failure. This can significantly improve
the efficiency of the strengthening design by increasing the
strain that can develop in the FRP. Hoop wrapping also
enhances the shear capacity of columns and prevents buckling of the axial fibres, enabling them to contribute in
compression. This contribution will be small since FRP
materials are weaker in compression than tension.
To calculate the required thickness of axial FRP, the effect of
hoop wrapping on compressive strength and strain to failure
of the concrete must be known. As discussed in Section 8.2,
the stressstrain behaviour of confined concrete is rather
different to that of unconfined concrete. It is perhaps worth
comparing the stressstrain response of the two types of
behaviour.
For unconfined concrete BS 8110 approximates the behaviour
as a parabola reaching a plateau at a value of 0.67fcu/mc with
strain terminating at 0.0035. However, while the initial
parabolic behaviour is reasonable, tests show that this is
normally followed by a descending stressstrain response,
shown by the solid line in Figure 31. The model in BS 8110
is a convenient approximation used for design purposes. For
confined concrete, the stressstain behaviour is again initially
parabolic, until sufficient concrete dilation occurs for the
hoop reinforcement to start confining the concrete. Provided
sufficiently stiff hoop FRP is used (i.e. the condition in
Equation 36 from Section 8.2 is met), this results in an
approximately linear ascending stressstrain branch, shown
by the dashed line in Figure 31. The additional confinement
also greatly increases the maximum achievable concrete
strain to failure (this can be taken conservatively as 0.01 and
is usually limited by rupture of the FRP). Equations 32 and
33 approximate this stressstrain behaviour, while Equations
39 and 40 correspond to the ultimate confined compressive
stress and strain, respectively, of the concrete. For a lightly
confined section (i.e. low stiffness hoop wrapping) the initial
parabolic stressstrain response may be followed by a branch
that initially descends, as shown by the dotted line in Figure
31. It is evident from Figure 31 that the three types of
behaviour are quite different. While it is not unreasonable to
assume a rectangular stress block approximation for unconfined concrete since the centre of pressure remains in approximately the same place, for the confined concrete this would
be conservative since, depending on the exact shape of the
stressstrain curve to failure, the centre of pressure of the
stress block may move considerably to the right, compared
with the relative shape of the unconfined stress block.

57

Design guidance for strengthening concrete structures using fibre composite materials

fcc

fcc
fccd

Fully confined

Fully confined for


x > D/2
1

Partially confined

fc0

Unconfined

Figure 31: Stressstrain behaviour of variably confined


concrete.

In axially loaded columns subject to moment, the confinement of the concrete varies. Rather than full mobilisation of
the confining hoop FRP due to dilation of the full crosssection, as would occur under concentric loading, when
moment is applied the dilation varies around the section in
proportion to the axial strain at any given point. Therefore,
the achievable level of confinement will vary. Fam et al.(145)
have proposed a variable confinement method where the
fully confined stressstrain relationship is reduced in
accordance with the eccentricity of the load (where the axial
force, N, and the moment, M, are related by the eccentricity,
e, i.e. e=M/N). While the general method seems rational,
although conservative, the procedure is rather complicated.
However, based upon their test results, there is little difference between using the fully confined stressstrain model
and a reduced confinement model provided that the position,
x, of the neutral axis from the outermost compression fibre is
greater than the radius of the column (i.e. the area of
concrete in compression is no less than half the total crosssectional area of the column):

Generally, the design of compression members strengthened


in flexure should consider the following:

58

at critical points the combination of maximum moment


and coexistent axial load
the risk of debonding
the risk of anchorage failure
the shear capacity of the column (see Section 8.3.4).

0.01

ccu

cc

Figure 32: Recommended stressstrain curves for combined


flexure and axial load.

It should be noted that flexural enhancement can only be


achieved within the span of the column. It cannot be achieved
at connections to beams or footings.
8.4.2 Moment capacity with axial load
To calculate the maximum moment and coexistent axial load
capacity of an FRP-strengthened column of circular crosssection, the following assumptions can be made:

(Equation 43)

If this condition is not met then a reduced stressstrain curve


should be used, referring to Fam et al.(145), or, conservatively, the unconfined concrete stressstrain model of BS
8110 or similar may be assumed. However, it is permissible
to extend the curve to a maximum ultimate strain of 0.01, as
shown in Figure 32. This is effectively the stressstrain
relationship of Equations 32 and 33 but with the slope of the
linear part equal to zero, i.e. E2=0. It should be noted that
this relationship is only applicable to columns of circular
cross-section.

Extended unconfined for


x < D/2 (i.e. E2 = 0)
Ec

cc

x > D/2

E2

Sections that are plane before bending remain plane after


bending.
Slip does not take place between the FRP and the concrete.
Axial fibres are placed in a layer of even thickness all
round the column.
The stressstrain response for unconfined concrete
follows the idealised curve for concrete presented in
current codes and standards, with mc=1.5
The maximum allowable compressive strain in the concrete is 0.01 or ccu, whichever is less.
The stressstrain response for steel reinforcement
follows the idealised curves presented in current codes
and standards, with ms=1.15. (As indicated in Section
6.2.3, many of the structures that require strengthening
will have been built before the 1997 edition of BS 8110.
Hence it would seem appropriate to use the earlier partial
safety factor of 1.15 rather than the value of 1.05 now
recommended.)
FRP has a linear elastic response to failure in tension.
The tensile strength of the concrete is ignored.
Confinement provided by any existing hoop steel is
ignored.
FRP in compression is conservatively neglected.

The procedure outlined below for calculating the required


thickness of axial FRP is based on the Highways Agency BD
84(10), Strengthening concrete bridge supports using fibre
reinforced plastics. The overall approach is, in principle, the
same as that used in conventional reinforced concrete column
design. However, because at the outset it is not clear whether
it is the (confined) compressive strain in the concrete or the
tensile strain in the FRP that is critical, the design procedure is
slightly more complex. The actual steps involved are given
below.

Strengthening axially loaded members

1. Select a thickness of axial FRP.


2. Using Equation 40, find the ultimate failure strain, ccu.
Set the strain in the outermost compression fibre of the
concrete to this value or 0.01 (to prevent excessive concrete crushing), whichever is less.
3. Assuming a linear strain distribution, perform sectional
analyses to determine by trial and error the depth, x, of
the neutral axis from the outermost compression fibre
when the axial load equals the applied ultimate load. If x
> D/2 then the force in the concrete in compression is
modelled using the fully confined stressstrain model of
Equations 32 and 33. If x < D/2 then the extended
unconfined stressstrain model should be used (Equations
32 and 33 with E2=0). The forces due to both the concrete
and FRP are most easily found using a layer-by-layer
approach, i.e. by subdividing the section into a number of
horizontal layers, parallel to the axis of bending, and
calculating the compressive force contribution of the
concrete (dependent on the average strain, and hence
average stress in the layer, from the stressstrain model)
and tensile force contribution of the FRP appropriately
for each layer. Tensile and compressive forces due to
existing steel reinforcement can be added separately.
4. Calculate the corresponding ultimate moment capacity.
5. If the ultimate moment capacity is less than the design
moment, M, increase the thickness of axial FRP and
repeat from Step 3 until the bending moment capacity
equals or exceeds M.
6. Evaluate the tensile strain in the FRP. If it is less than the
ultimate design strain then the axial FRP thickness is
adequate. However, if the strain in the FRP exceeds the
ultimate strain capacity, the FRP will fail before M is
reached. This implies that it is the strain in the FRP and
not the concrete, as assumed in Step 2, that is critical.
Therefore, repeat Steps 3 to 5 but this time limit the strain
in the FRP to its ultimate design value. Obviously, the
maximum strain in the concrete will be less than ccu but
the same stressstrain model may be used.
7. In order to ensure ductility, check the maximum compressive strain in the concrete. Provided it is not less than
0.0035, the axial FRP thickness is adequate. Otherwise,
increase the strain in the outermost compression fibre of
the concrete to 0.0035 and proceed to step 8.
8. Increase the axial fibre thickness and determine by trial
and error the depth of the neutral axis when the axial load
equals the applied ultimate load. Determine the corresponding ultimate moment capacity.
9. Evaluate the tensile strain in the FRP. If it exceeds the ultimate capacity, repeat step 8 until the tensile strain in the
axial FRP is equal to or less than its ultimate strain limit.
This design procedure has been set up so that the strain in the
concrete does not fall below 0.0035 in order to ensure
ductility under dynamic loading. However, lower concrete
strains may be acceptable where flexural strengthening is
required for reasons other than dynamic loading.

Use of this procedure to determine the thickness of axial


FRP required to strengthen columns in flexure is very
tedious. For convenience, therefore, design charts have been
developed by Denton et al.(146). The design charts form part
of the Highway Agency Design Document BD 84(10), in
which their application is explained and examples are
presented. They are based on the BS 5400: Part 4 models for
concrete and steel reinforcement. Also, any contribution to
the flexural strength of the column by the FRP in compression has been neglected. Note that the confined concrete
stressstrain model assumptions are not the same as those
described in the above procedure, but are conservative.
8.4.3 Debonding
The work by Cuninghame et al.(147) appears to show that,
providing the number of axial layers is not excessive, the
provision of hoop wrapping over the axial fibres can prevent
debonding failure. However, debonding failure may occur in
wrapped columns (below the load capacity expected on the
basis of the area of FRP) and it is important therefore that a
degree of caution is exercised until there is more reliable
information on this aspect of design.
8.4.4 Anchorage
The axial FRP can fail at the base and/or top of columns and
at cut-off points. Anchoring the FRP by extending beyond
the point at which it is theoretically no longer required can
prevent this. The anchorage length can be determined using
the principles described in Section 6.3. In most situations,
for example where a column is built into a rigid base, it is not
feasible to provide sufficient anchorage length. In such
situations, the axial wrapping may be enclosed within a
collar constructed of steel or concrete.

8.5 DUCTILITY
Lack of ductility is largely an issue for compression
members that are located in seismic regions. Upgrading
normally involves confining the concrete at column ends
(where bending moments are greatest) with hoop FRP. To
ensure that column bar buckling does not control the flexural
failure mode, additional checks on the transverse reinforcement ratio need to be performed, particularly for slender
columns where M/VD > 4, in which M and V are the
maximum column moment and shear, respectively (see
Priestley et al.(101)). Ductility enhancement may increase the
risk of shear failure both at column ends and column centres,
and the risk of flexural failure due to lap splice debonding at
the junction between the footing and column base(102).
As noted in Section 5.7.2, seismic loading is not a major
loading case for most UK structures and will therefore not be
discussed further.

59

Design guidance for strengthening concrete structures using fibre composite materials

8.6 STRENGTHENING COLUMNS WITH


NON-CIRCULAR CROSS-SECTION

Confining square columns is generally acknowledged to be


less efficient than confining circular columns. This is largely
attributable to the fact that, with square columns, confinement is concentrated at the corners rather than over the entire
perimeter. Current research on small columns has shown that
the maximum achievable increase in compressive stress for
FRP-confined square columns with reasonable levels of
rounding of corners is about 50%, compared with up to
200% for circular columns. For larger columns, found more
often in practice, the level of increase may be less than this.
The efficiency decreases further with columns of rectangular
cross-section and/or large side dimensions. (The difference
in confinement performance between circular and rectangular
columns is similar to the difference between the use of hoops
and stirrups (without cross-ties) in conventional steelreinforced column design).

Several models of the strength enhancement of rectangular


columns have been proposed and compared with the limited
number of experimental results available. However, it should
be noted that most models are semi-empirical in nature and
have been calibrated with small-scale test specimens.
Typically these specimens are 150mm x 150mm for square
columns and up to 150 mm x 225mm (i.e. an aspect ratio of
1.5) for rectangular columns. As rectangular columns get
larger in size the length of the unconfined regions along the
sides of the column increases creating a size effect, which is
not evident for circular columns. It should also be noted that
the confinement model is based upon concentrically loaded
column behaviour. It is likely that if there is significant load
eccentricity then the stress distribution will be such that the
effectively unconfined areas will be more highly stressed
than assumed in the model. Caution should be used in such
a situation.
The generally accepted theoretical approach is to develop an
area of effective confinement defined by four parabolas
within which the concrete is fully confined and outside of
which negligible confinement occurs. The shape of the parabolas and the resulting effective confinement area is a function of the dimensions of the column and the radius of the
corners. From this effective area a shape factor can be defined.
An effective confining pressure must be calculated based
upon the same thickness FRP wrap confining an equivalent
circular column of diameter D. The effective strength of the
confined concrete is then given by the unconfined strength
plus the product of the shape factor, the effective confining
pressure and an effectiveness coefficient, which is derived
empirically and conservatively taken as 2.0. The various
models differ mainly in the calculation of the effective area
and the diameter of the equivalent circular column. Use of
any current state-of-the-art explicit method for analysing
strengthened square or rectangular columns is not
recommended unless the following conditions are met:

60

Loading is essentially concentric.


The smaller edge dimension is no greater than about
200mm.
The aspect ratio is no greater than 1:1.5.
The corners can be provided with a radius of at least
15mm to avoid stress concentrations.

Of the various models suggested, that by Lam and Teng(148)


predicts the observed behaviour most accurately. However,
while their model suggests that the initial slopes of the
parabolas should be the same as the slope of the diagonal
lines between the column corners, a more common assumption(149) is that the initial slopes of parabolas start at 45 to
the face of the column, as shown in Figure 33. Although this
assumption may be conservative, it is recommended that
such an assumption be made in calculating the confined area.
Thus, the ratio of the effective area of confinement
(contained within the parabolas) and the total cross-sectional
area of the column is given by:

Ae 1 (h 2 Rc ) + (b 2 Rc ) 3 Aol / (3 Ag ) sc
(Equation 44)
=
Ag
1 sc
2

where
sc =
b =
h =
Rc =
Ae =
Ag =

is the ratio of longitudinal steel to the cross-section


length of short side
length of long side
radius of corner
effectively confined area
total cross-sectional area

=
Aol =
=

bh (4-)Rc2
area of overlap of the parabolas
0
4lol3
+ lol (2b (h 2 Rc ) )
3(h 2 Rc )

=
where
lol =

if 2b >
(h 2Rc)
if 2b < (h 2Rc)

length of overlapping region

(h 2 Rc ) 2 b(h 2 Rc )

4
2

effectively
confined
region, Ae

Rc
45o
h

Figure 33: Assumed confined region for FRP-wrapped


rectangular column.

Strengthening axially loaded members

Subtraction of the area Aol is required since, as the aspect ratio


of the cross-section increases such that the limit 2b < (h 2Rc)
is reached, the larger parabolas corresponding to the longer
edge will overlap in the middle of the section, as shown in
Figure 34. This resulting area is not confined and therefore
must be subtracted from the total area enclosed by the
parabolas.

lol

effectively
confined
region, Ae

b
Aol

45o

Confinement efficiency can be improved by rounding the


corners of the column or by casting circular or oval concrete
rings around the column perimeter. However, it should be
noted that the confined strength of oval columns reduces as
the aspect ratio increases. Teng and Lam(150) suggest using
Equation 47 but with an empirically derived shape factor:

c
gs =
a

where a and c are the major and minor dimensions,


respectively.

Figure 34: Overlapping parabolas in confined region.

fr =

Thus, a shape factor is defined as:

b Ae
h Ag

(Equation 45)

The aspect ratio in this equation is necessary in order to reflect


properly the reduction in confinement as the aspect ratio
increases. The shape factor relates the effective confining
pressure to that provided by an FRP wrap of the same
thickness to an equivalent circular column of diameter D.
The diameter of the equivalent circular column is defined by
Lam and Teng as the diagonal distance across the section, i.e.
D=(b2+h2). Thus, using Equation 38, the equivalent
confining pressure is given by:
fr =

2 f fd tf

(Equation 46)

b2 + h2

Given the equivalent confining pressure and the shape


factor, the strength of the confined rectangular column is
given by:
f cc = f c0 + 2.0 g s f r

where
fc0 =

(Equation 48)

The equivalent circular column is defined as one with the


same volumetric FRP ratio as the elliptical column and, therefore, the equivalent lateral confining pressure is given by:

gs =

(Equation 47)

[1.5(a + c)
2ac

ac f fd tf

(Equation 49)

Again, this model should be used with caution, since it is


calibrated against a small number of concentrically loaded
specimens. The greatest aspect ratio tested was 5/2.
Rectangular carbon fibre jackets are also reported to provide
sufficient confinement and bar buckling restraint to achieve
high flexural displacement ductility levels(102). Perhaps an
alternative approach may be to use preformed circular FRP
shells and fill the intervening space with grout. Other methods
may also be employed in order to improve the efficiency of
confined rectangular columns. Where it is difficult to shape
corners to a radius of at least 15mm significant stress
concentrations may occur resulting in premature failure of
the FRP. In such situations it has been suggested that additional localised wrap reinforcement be provided at the corners
prior to the application of the continuous layers, thus reducing corner stresses(151). It has been suggested that internal
FRP ties can be used through the width of the section to
increase confinement where side lengths are large(152). It
may be beneficial to incorporate additional longitudinal FRP
in a wrapping scheme to reduce the likelihood of bursting
failures. The design of strengthening schemes that use any of
these systems should be verified by independent testing.

0.67 fcu / mc
To summarise the information currently available:

Again, it should be emphasised that this model has only been


calibrated against small-scale concentrically loaded columns
and caution should be taken if applied to columns not typical
of those tested in the literature due to the size effect of the
unconfined sides resulting in premature failure.
Shear strengthening of rectangular columns should be
carried out according the methods outlined in Chapter 7, in
the same way as for beams.

A significant increase in the axial load capacity of square


or rectangular columns by wrapping with FRP may be
difficult to achieve in practice.
An increase in the flexural strength of square columns by
wrapping with axial and hoop FRP should be feasible,
but any proposed design should be verified by independent testing.
Ductility enhancement of square columns may prove
difficult.
FRP can be used to increase the shear capacity of square
and rectangular columns.

61

Design guidance for strengthening concrete structures using fibre composite materials

62

Strengthening axially loaded members

63

Design guidance for strengthening concrete structures using fibre composite materials

64

EMERGING TECHNOLOGIES

9.1 EMERGING STRENGTHENING


TECHNOLOGIES ALREADY USED IN
PRACTICE
9.1.1 Prestressing using FRP composites
The use of prestressed FRP composites for flexural strengthening of concrete structures has been developing over recent
years, and a number of proprietary FRP systems are now
available commercially. However, design guidance in this
area is still developing, and as such it is advisable to seek
specialist advice for the design of prestressed FRP composite
strengthening systems.
There are many reasons for strengthening of concrete structures using prestressed FRP composites:

increasing live load capacity


reducing dead load deflections (i.e. mobilising locked in
stresses)
reducing crack widths and delaying the onset of cracking
reducing serviceability problems such as excessive deflection, cracking of the concrete and tensile steel stresses at
serviceability
improving fatigue strength by reducing tensile steel
stresses
regaining prestressed conditions in the concrete that may
be lost by damage to the original prestressing tendons or
other effects.

However, if the concrete stress at the serviceability limit state


is high, prestressing with FRP composites will not provide a
significant increase in the serviceability load.
For some prestressing systems, no reliance on an adhesive
bond is required, which can be advantageous where very low
or very high service temperatures are present (or there is a
significant fire risk) and could significantly reduce the performance of the adhesive, or where the surface quality of the
concrete is inadequate for adhesive bonding.
The use of prestressed FRP composite strengthening systems
may not be appropriate, or at least will require more detailed
planning and risk assessment, where there are limited available installation periods (for example within railway possessions or in structures where there is a continuous industrial
process), significant risk of vandalism, or exposure to a highly
aggressive environment that would affect the long-term
properties of the FRP composite.

It has also been suggested that the use of prestressed FRP


composites can increase shear capacity by a confining effect
on the concrete(153), although this has not been investigated
in great detail.
FRP composites generally exhibit superior durability and
fatigue properties to those of steel. A number of types of FRP
composite can be used to post-tension existing concrete
structures: bonded or unbonded FRP composite plates or
sheets, bonded FRP NSM reinforcement and external FRP
composite tendons.
Prestressing allows a greater proportion of the FRP composite tensile strength to be utilised, and can therefore be more
efficient than unstressed solutions. Carbon FRP composite is
generally the most suitable type for prestressing applications
due to its superior creep and stress rupture properties
compared with those of glass FRP and aramid FRP
composite. It is advisable not to use glass FRP composites
for prestressing, unless the prestressing force being applied
is quite low, where stress rupture will not be an issue.
Aramid FRP composites can be used instead of carbon FRP
composites, their lower tensile elastic modulus being an
advantage in achieving greater control of elongations in the
prestressing process. However, as with glass FRP, aramid
FRP can be susceptible to stress rupture, and therefore the
prestress levels should be limited.
The basic principle of strengthening concrete structures by
post-tensioning with FRP composites is similar to that for
conventional post-tensioned concrete structures. A portion or
all of the existing dead load in the concrete member is
transferred to the FRP composite by creating a tensile force
in the FRP prior to application, bonding and/or anchoring the
FRP to the concrete substrate, and then releasing the prestress load. The bonded joint or mechanical anchors at the
ends of the plate then transfer the prestress into the concrete
member. The prestressed FRP composite carries both a
portion of dead load, and also live load, in comparison to
unstressed FRP strengthening where the FRP composite
carries only a portion of live load.
The method of prestressing the FRP composite is crucial to the
feasibility of a practical FRP prestressed strengthening application. The method can be described in a number of stages:
1. application of tensile force to FRP composite
2. anchorage/bonding of FRP composite to concrete substrate
3. release of prestress into concrete member and redistribution of forces throughout section.

65

Design guidance for strengthening concrete structures using fibre composite materials

The method of jacking out the dead-load deflection of a


concrete member, or physically removing existing dead load
temporarily, prior to the application of unstressed FRP
composite is also essentially a prestressing solution as the
FRP composite carries a portion of the dead load in addition
to live load.
The ROBUST project demonstrated the benefits of prestressing the FRP prior to bonding to the concrete, on 1.0m
and 4.5m beams in the laboratory, and with 18.0m beams in
the field. Specially developed glass FRP end tabs were
developed to enable the carbon FRP plate to be pulled, prior
to anchoring the tabs into the concrete using resin
anchors(29).
Anchorage of the prestressed FRP composite is a critical
aspect of the application. Large shear stresses are present
within the adhesive bond when the prestress in the FRP
composite is released and transferred into the concrete
member. Therefore, in most cases a mechanical anchorage at
the ends of the FRP composite is required, although
techniques are under development to avoid the need for
bulky end anchorages.
A number of anchorage types have been developed,
generally based on jacking the FRP composite plate through
a steel anchorage and bolting through the plate (which is
locally strengthened with steel or glass FRP end tabs to
avoid failures due to stress concentrations around the bolt
holes) into the anchorage and the concrete substrate. The
plate is then cut and the transfer of prestress occurs via the
bolted anchorage.
In Switzerland, a system has been developed whereby a
carbon FRP plate is stretched over a set distance between
two large wheels(154). The entire mechanical system is then
lifted up to the soffit of the concrete and the laminate is
bonded to the structure. This method overcomes many of the
problems associated with successfully gripping the plate.
Furthermore, this method also allows the problem of high
longitudinal shear stress at the concreteplate interface to be
overcome. The longitudinal shear stresses originally led to
anchorage failures in laboratory tests. However, a graduallyanchored system was devised to reduce the longitudinal
shear stresses and delay the onset of anchorage failure.
Using this method, the plate is bonded to the concrete from
the centre and then moved outwards in stages. As each
portion of the plate is bonded to the concrete at successive
stages, the prestressing force is slightly reduced to a nominal
value at the end of the plate. In order to speed up the curing
process, so that the step-wise technique is economic and
practical, heating devices within each portion of the plate are
used to reduce the adhesive bond curing time.

The relaxation loss for prestressed FRP composite in comparison to high strength steel is generally small.
The losses in prestress in the long term are essentially the
same as those for post-tensioning with low-relaxation steel,
due to the reduction in elastic modulus of the concrete and
shrinkage in the long term.
The ultimate load capacity of concrete members posttensioned with FRP composites can be analysed based on
conventional theory for reinforced-concrete structures(29,155),
but only if flexural failure is the dominant failure mode. The
failure mode may be either concrete crushing or FRP
composite rupture, depending on the degree of prestress
applied to the FRP.
Design checks are also required at the serviceability limit
state. The level of prestress should be such that the following
conditions are acceptable:

66

tensile stress and cracking at the concrete edge away


from the prestressed FRP composite under dead and
superimposed dead load, after transfer of the prestress
compressive stress in the concrete and dead, superimposed dead and live loading
tensile stress in the FRP composite under dead and superimposed dead loads (for durability and stress rupture),
and live load (for fatigue).

The achievable prestress levels vary depending on the type


of FRP composite used, based on their susceptibility to stress
rupture. Carbon FRP composite is not susceptible to stress
rupture and therefore the greatest degree of prestress can be
achieved, followed by aramid FRP composite and glass FRP
composite, which both exhibit stress rupture. As an indication, carbon FRP composite can typically be stressed up to
50% of the design tensile stress, aramid FRP composite up
to 30% of the design tensile stress and glass FRP composite
to only 1520% of the design tensile stress. For prestressing
systems where the adhesive bond is relied upon to transfer
prestress forces, the permanent stress in the adhesive should
be limited to 25% of the design strength to avoid creep and
durability problems, as stated in Section 5.6.8.
A number of commercial prestressing systems are currently
available, using the following methods of prestressing:

On transfer of the prestress to the concrete member, some


losses occur in a similar manner to those for conventional
prestressed post-tensioned structures. The losses in prestress
in the short term are due to:

elastic shortening of the concrete member


creep effects in adhesive for prestress of systems with no
mechanical anchorage
drawing in within the anchorage system (if a mechanical anchorage is used).

Stressing and anchoring of carbon FRP plates in a steel


anchorage, placed in a recess in the concrete, containing
a base plate for force transfer (bonded and bolted to the
concrete), tensioning plate for the hydraulic jack and
levelling aids: the stressing process is undertaken in two
stages with temporary and permanent anchorages.

Emerging technologies

Stressing and anchoring in a steel and carbon FRP


composite anchor block, placed in a recess in the concrete substrate: the steel anchorage is bonded and bolted
to the concrete substrate; the prestressing operation is
carried out in a single stage.

Some prestressing systems are designed such that plate


failure would always occur prior to anchorage failure.
A number of reinforced concrete and conventionally prestressed and post-tensioned concrete structures throughout
Europe, particularly in Germany and Switzerland, have been
strengthened using prestressed FRP composites. The first
full-scale application of a FRP composite prestressing system
in the field was on Lauterbridge, Gomadingen, Germany in
October 1998. In this particular case, the prestressing system
was installed to reduce crack widths, and increase the
flexural strength and rigidity.
A number of existing concrete bridges have been strengthened using carbon FRP prestressing tendons or aramid FRP
ropes(156,157). FRP tendons were chosen due to easy handling
on site, small-diameter bundles allowing simple deviator
detailing, and corrosion resistance. Anchorage systems for
FRP tendons have been developed that allow full design
strength of the tendon to be assumed, under both static and
cyclic loading conditions(158). Furthermore, research has
shown that external FRP tendon prestressing systems allow
substantial rotation to occur over supports in continuous
spans near the ultimate limit state. Based on this research, it
is possible for full moment redistribution to be assumed(159).
Examples of commercial prestressing systems are shown in
Figures 35 and 36.
9.1.2 Unstressed FRP anchorage techniques
The use of transverse FRP U-wraps around the soffit of a
beam is a common laboratory technique to anchor longitudinal FRP laminates(160). This technique has also been carried
out in practice on many occasions, with satisfactory results
being achieved(161).

Figure 36: Prestressing system used on Lauterbridge,


Gomadingen.

Another well-founded laboratory technique to anchor wet


lay-up FRP laminates is to cut a transverse groove in the
concrete, insert the end portion of the fabric into the groove
and anchor into place (through resin) an FRP NSM bar(127).
Again, this technique has also been used in practice, particularly where sufficient anchorage length for the FRP cannot
be provided by bond alone. Examples where this technique
is particularly useful include anchorage of U-wrap laminates
for shear strengthening of T-beams(114) and anchorage of
longitudinal laminates for flexural strengthening of members,
which vary in cross-section near supports(54).
9.1.3 Bolted plate anchors
Nurchi et al.(162) have assessed the enhancement of anchorage provided by bolting the ends of an FRP plate to concrete,
in addition to adhesive. The primary use of the bolts is to
prevent or delay the onset of debonding and concrete cover
failure. However, the use of bolts requires multi-directional
FRP in order to prevent longitudinal splitting failure of the
FRP, and in order to provide sufficient bearing stiffness at
the position of the bolts. The bolts, about 200mm in length,
are fixed into predrilled holes using epoxy adhesive. An
additional FRP reinforcement layer is added at the ends of
the FRP plates to increase strength at these positions.

Figure 35: Dead-end prestressing system (left) and stressing anchorage (right).

67

Design guidance for strengthening concrete structures using fibre composite materials

The results of their tests indicate that the use of bolts within
the shear span of the beam significantly postpones debonding.
Following debonding, they are still capable of anchoring the
ends of the FRP so that it acts as an unbonded tension member,
resulting in a less brittle mode of failure at the ultimate
condition. The load capacity of the beam following debonding
is reported to be at least equal to the capacity of the beam
with the bond intact, although greater deflections occur.

resistance of concrete beams considerably. In particular, de


Lorenzis and Nanni(166) have shown that, in the case of
strengthening T-beams in shear, if it is possible to anchor the
NSM bars into the compression flange, shear resistance is
enhanced greatly. They have also shown that angling the
NSM bars at approximately 45 (rather than placing them
vertically) is advantageous. While such angled reinforcement is problematic for plates or sheets, it is relatively
straightforward for the NSM case.

9.1.4 Concrete masonry walls


9.2.3 FRP anchor systems
FRP composite systems can provide solutions for the strengthening of masonry. The dynamic properties of the existing
structure remain unchanged because there is minimal
addition of weight, and stiffness changes may be engineered
case by case.
FRP composites used for flexural and/or shear strengthening
of masonry structures are similar to those used for strengthening of concrete elements. A number of research projects
have demonstrated the effectiveness of FRP composites to
improve the structural performance of masonry walls, particularly in situations of high slenderness(163). Available
literature shows that walls strengthened with FRP in the
laboratory usually fail due to debonding of the laminates.
An alternative to FRP laminates is the use of NSM FRP bars.
In this way, FRP bars have been used successfully to increase
the flexural strength of concrete masonry walls. FRP systems
can also be used to improve the performance (strength or
pseudo-ductility) of masonry walls subject to in-plane
loads(164). Strengthening by FRP structural repointing (insertion of small diameter FRP bars in the bed joints) can also
increase shear capacity and provide pseudo-ductility to walls.

9.2 EMERGING STRENGTHENING


TECHNOLOGIES AT THE RESEARCH
STAGE

In situations where FRP laminates need to be mechanically


anchored, due to insufficient bond length being available,
there are various options already available. For instance,
bolted steel plate anchor systems, as described in Section
9.1.3, may be used for FRP plates. Alternatively, FRP NSM
bars may be used to anchor wet lay-up FRP laminates within
slots(127), as described in Section 9.1.2.
Another anchorage system in the process of development
has proved to be very effective in preliminary tests at the
University of Missouri-Rolla(114). The anchors themselves
consist of many glass fibres of overall length 150mm. They
are dipped into resin to a depth of 75mm, and the resulting
glass FRP portion (of approximate diameter 12mm) is allowed
to cure. Holes 75mm deep and 200mm apart are drilled into
the soffit of the concrete structure to be strengthened, and
resin is inserted into each hole. The first layer of wet lay-up
sheet is adhered to the surface of the concrete, following
standard procedures. At each hole location, the glass FRP
end of each anchor is inserted into the hole through the sheet
(by locally realigning the sheet fibres to skirt around the
anchor). The dry end of each anchor is then fanned out over
the surface of the first layer of sheet and fixed into place with
resin. Subsequent layers of sheet are added above the fanned
anchors so that, after curing, the fan anchor is located within
the thickness of the FRP laminate. This enhances anchorage
and bond behaviour.

9.2.1 Prestressed NSM bars


9.2.4 Steel-reinforced polymers
One way in which the problem of high shear stress at the
ends of prestressed FRP plates might be overcome is to use
NSM bars instead. This is because they are bonded over
most of their perimeter, leading to a spreading of longitudinal shear stress in comparison with the one-side-bonded
laminate case. Research conducted in Sweden and in the
USA has confirmed their suitability(165). Naturally, the main
concern is how to prestress the NSM bars prior to insertion
into the grooves. At present, this is being considered in some
detail because of the potential for this form of strengthening.
9.2.2 NSM bars for shear strengthening

The use of steel-reinforced polymer (SRP) materials has


recently been considered in the USA(167). Part of the motivation for use of this novel material is that discarded tyres
contain significant quantities of Hardwire steel strand,
which can be used to make the SRP. This clearly has
environmental benefits, which is advantageous. However,
the thin steel strands are susceptible to corrosion. Various
matrix materials have been looked at to protect and bind the
steel strands. A cementitious grout presently appears to offer
a good compromise between structural strength and a
durable composite.

Because NSM bars offer potentially superior bond performance over plates or sheets, their use for shear strengthening
has been attempted, with success. It has been found that the
use of either glass or carbon FRP NSM bars increases shear

A parking garage in Indiana, USA, was recently condemned.


Prior to its demolition, parts of it were strengthened using
SRP, with encouraging results. The strengthened sections
showed distinct improvement over the original sections in

68

Emerging technologies

terms of capacity and ductility. Failure occurred by SRP


peeling, in much the same way that might be expected to
occur in the FRP situation.
9.2.5 Deep embedded bar for shear strengthening
Particularly when working on concrete bridges, it is desirable that any strengthening be carried out rapidly, with as little
disruption as possible. In order to expedite shear strengthening, it is possible that vertical holes could be drilled into
the bridge from soffit level, and FRP bars fixed into the holes
with resin to provide additional shear resistance(168).
Although similar in concept to the use of stainless steel
threaded bar for such applications, the FRP alternative
requires access only from the soffit (rather than from top and
bottom), and the possibility of corrosion of reinforcement in
close proximity to the stainless steel bars is eliminated. This
technique has the further benefits that no surface preparation
of beam webs is required in beam-and-slab situations and the
webs need not be accessible. This is very useful in situations
where precast beams are closely spaced.
This shear strengthening technique has been shown to be
feasible, preventing shear failure in laboratory tests and
instead forcing a ductile flexural failure to occur.
9.2.6 Prestressed carbon FRP straps for shear
strengthening
As it is difficult to anchor U-wraps around T-beam webs,
slots could be drilled in the flange at regular intervals, and
full strap-wrapping of the equivalent rectangular section
carried out. Since local overstrain in bonded FRP at crack
locations can lead to local failure, it would be best not to
bond these straps to the concrete. Furthermore, in order to
resist shear crack openings (and hence enhance aggregate
interlock effects), the straps should be prestressed.
Such a system has been developed for commercial use(169).
The straps consist of five or ten individual carbon FRP tapes,
heat-welded together in order to create continuity. The straps
are prestressed using a patented system, and accuratelymachined wedges are inserted between strap and concrete in
order to maintain the prestress. Results from tests show that
shear strength is enhanced considerably through making use
of a debonded, prestressed system.
9.2.7 Mechanical fastening techniques
Some of the limitations associated with FRP strengthening
of concrete structures are the modest longitudinal shear
strength of adhesives, the time taken to prepare the concrete
surface and the brittle form of failure, which occurs when the
FRP peels off under severe overload. The use of a mechanical bonding system might be an approach to addressing
these problems.

Research carried out in the USA has shown the feasibility of


using such a technique. Originally intended for extremely
rapid strengthening of civilian concrete bridges to allow heavy
military vehicles to pass, the technique is to mechanically
fasten a precured FRP laminate to the concrete soffit using
power-driven bolts. No surface preparation of the soffit is
required. The FRP is bidirectional in nature, to prevent
longitudinal splitting. Preliminary tests carried out in
Madison, Wisconsin, have shown the potential for this rapid
form of construction(170), although integrity of the cover
concrete following the power-driving activities has been
shown to be crucial to successful implementation.
9.2.8 Strengthening for torsion
Limited tests have been conducted on torsional strengthening of concrete structures using FRP materials. Published
results show that such strengthening is possible and that FRP
can contribute substantial torsional resistance(171). Preliminary indications are that the wrapping of the torsion element
should be as full as possible, as the confinement that is
created in this way is particularly beneficial in resisting
torsion in a controlled, ductile manner.
9.2.9 Life expectancy modelling
Research carried out in Italy is attempting to rationalise the
life expectancy of FRP-strengthened concrete schemes(172).
In order to do this, a series of degradation states has been
defined, each component relating to a particular intervention.
These interventions are, in order of least to most severe:

inspection
cleaning
repair
partial substitution
total substitution.

A database is being developed on a range of FRP-strengthening schemes throughout the world. By categorising performance levels and long-term mechanical properties of the
constituent materials, it is hoped that this research will lead
to a rational approach to life-expectancy predictions, one of
the most important pieces of information in any design.
9.2.10 Inorganic adhesives
Because of concerns over the performance of organic
adhesives at elevated temperatures, there are moves to
develop inorganic adhesives, more akin to cement-based
materials. Toutanji and Deng(173) have reported tests on beams
strengthened using Geopolymer adhesives consisting of
alumino-silicate with a water-based activator. These perform
significantly better at high temperatures and, being waterbased, are somewhat easier to use on site.

69

10 WORKMANSHIP AND
INSTALLATION

10.1 INTRODUCTION
The design guidance in Chapters 5 to 8 is only valid if the
component materials used are in accordance with the specification and the installation is carried out correctly. This
Chapter is not intended to be a specification for strengthening with composites but gives background information on
the standards of workmanship and the installation procedures required. Further information on the requirements for
inspection during the installation process, along with advice
on the necessary records that need to be maintained, is given
in Chapter 4 of TR57(5), which should be read in parallel
with this Chapter.
The Client should be satisfied as to the competency of the
contractor. All installations should comply with the requirements of the Health and Safety at Work etc Act(174), the
Control of Substances Hazardous to Health Regulations(27)
and the Construction (Design and Management) Regulations(26). In addition, all materials must be used in accordance with the manufacturers requirements.
Only limited post-application inspection is possible, so the
success of the application relies heavily on the quality of the
workmanship. It is therefore crucial to the success of the
installation that an experienced contractor, with suitably
trained and supervised staff experienced in the technique, is
appointed. The contractor should have quality assurance
procedures in place, accredited and audited in accordance
with ISO 9002(175). In the UK, the contractor should have a
proven track record in the installation of composites and
preferably be a member of the Concrete Repair Association.
The contractor must be able to demonstrate competency and
be approved for the application of the system. This approval
may be obtained by providing evidence of the training of the
operatives who will undertake the work and by documentary
evidence of experience on similar projects.
It is strongly recommended that the following issues be
taken into account when selecting a contractor:

The contractor should provide a full method statement


and risk assessment for the works.
Operatives should be trained and qualified in application
techniques by the manufacturer of the system.
Personnel should be supplied with the correct personal
protection equipment for use when handling the materials.

The contractor should provide a safe means of access to


the work location and maintain an environment suitable
for the successful use of structural adhesives.
Procedures should be in place to minimise the risks to the
workforce and to any other persons (especially children)
who may be affected by the work.

The following sections give general guidance on the


installation of plate and fabric materials, which are bonded
to the surface of the concrete, and of NSM material. The
installation of shells around columns, which are generally
bonded to the concrete by means of a secondary process such
as grout injection, should be in accordance with the
manufacturers requirements and are not covered here. For
all materials and processes, quality assurance procedures
should ensure that each stage is approved before starting the
next stage. It is vitally important that the manufacturers
recommendations are followed throughout.
The sequence of the subsections in this Chapter follows the
step-by-step procedures that would be followed on site by a
competent contractor.

10.2 EVALUATION OF CONCRETE


CONDITION
An investigation of the condition of the structure(7) should be
carried out prior to the decision to undertake strengthening.
This will identify any deterioration processes (e.g. reinforcement corrosion due to the presence of chlorides) likely
to affect the performance of the structure within its residual
design life. The investigation should also include a thorough
inspection of the concrete surfaces on which the bonding is
to be carried out, a visual inspection, an assessment of the
concrete strength (see Figure 37) to assess whether it is sufficient for strengthening to be carried out (see Section 2.2),
chemical analysis and a sounding survey to identify defect.
If defects are identified, repairs should be carried out using
an appropriate concrete repair system in accordance with the
manufacturers recommendations. Cementitious repairs
should be cured for at least 28 days before undertaking
bonding work. As indicated in Section 2.2, the cause of any
deterioration should, as far as possible, be eliminated before
the structure is strengthened.

71

Design guidance for strengthening concrete structures using fibre composite materials

steam cleaning alone or in conjunction with detergents


for smaller areas, mechanical wire brushing or surface
grinding (see Figure 38).

Mechanical impact methods such as needle gunning and


bush hammering are very effective but are often too aggressive and produce a deeper texture in originally smooth
concrete. In addition they may shatter aggregate particles
causing micro-cracks.

Figure 37: Use of pull-out test to determine concrete strength.

10.3 CONCRETE PREPARATION


10.3.1 Concrete surface for plates and fabric
Good preparation of the concrete surface is of paramount
importance to the long-term success of the bonding and
strengthening operation, though this is less crucial for fully
wrapped systems. Before adhesive is applied the concrete
surface must be cleaned so that it is free of laitance, loose
material, fungal or mould growth, oil or grease, corrosion
products, previous coatings and, in the case of new concrete,
mould release agents and curing membranes.
It is important that the preparation process selected is such
that it removes the surface layer to expose small particles of
aggregate without causing micro-cracks or other damage in
the substrate. The surface should not be polished or roughened excessively. Sharp edges, shutter marks or other irregularities should be removed to achieve a flat surface.

Mechanical methods may be effective in removing deeply


penetrated greases, oils and paints, but may remove an unacceptable depth of concrete. Washing techniques may be
ineffective and can simply spread the contaminant further. In
such instances the use of solvent-based and sodium
hydroxide-based products in the form of a gel or poultice can
be effective in drawing out the contaminants. Such products
must be used with great care; if they are not thoroughly
removed, debonding of the strengthening system may occur.
Wet grit-blasting or vacuum dry-blasting are commonly used
because they reduce the dust created by open dry-blasting
which is unacceptable for health, safety and environmental
reasons. However, wet techniques may create a water disposal
problem and the concrete surface needs to be allowed to dry
out to a degree that is suited to the intended adhesive.
Whichever method is adopted it is always advisable to carry
out trials to select and optimise the technique in conjunction
with the material supplier.
The preparation of the surface should be to a standard such
that the adhesive layer is of uniform thickness when the
strengthening material is in place. Any steps in the surface
should be removed and hollows filled with a suitable quicksetting repair mortar (see Figure 39). Generally the flatness
of the surface should be such that the gap under a 1m
straight-edge does not exceed 5mm. The thickness of the
adhesive layer is commonly between 1 and 5mm depending
on the material, though thickening to 10mm may be acceptable to accommodate local defects such as dislodged
aggregate.

Figure 38: Surface grinding.

Effective preparation techniques include:

72

wet-, dry- and vacuum-abrasive blasting


high-pressure washing, with or without emulsifying
detergents, and using biocides (where necessary)

Figure 39: Filling imperfections with quick-setting repair


mortar.

Workmanship and installation

When fabric is to be wrapped round corners, e.g. round a


square column or round the bottom of a beam, the corners
should be rounded to a minimum radius of 15mm, or as
recommended by the supplier, to avoid local damage to the
fibres.
Minor imperfections in the concrete surface can be treated at
this stage with epoxy materials which can be applied in thin
layers and whose rapid strength gain permits over-bonding
to be carried out after a short time. Some bonding systems
require the use of a primer on completion of the surface
preparation. This primer, which seals the surface, should be
applied in strict accordance with the manufacturers
instructions.
The final assessment for surface quality can take the form of
a series of pull-off tests. (If a surface primer is used, the tests
should be carried out on the primed surface.) Figure 40
shows a dolly after being pulled off, with the concrete still
adhering to it. A minimum of three tests per representative
area should be carried out, as described in BS 1881: Part
207(176), to give an indication of the tensile strength of the
substrate and the quality of the surface preparation. The concrete surface should be dry for normal applications. Where
this is not possible, because of the nature of the structure,
special consideration should be given to the adhesive to be
employed.

10.4 MATERIAL CONFORMITY


The design process will involve certain assumptions about
the properties of the FRP. It is therefore important that all
materials are in accordance with the specification. FRP
plates, rolls of fabric, etc., should carry identification labels
to indicate their type and grade. To ensure that materials are
compatible, they are generally specified as part of a system,
e.g. pultruded plate and adhesive. Other materials should not
be substituted without the approval of the specifier.
Guidance on the approaches for ensuring conformity and on
acceptance tests are given in TR57(5), which includes a
proforma for recording details of the materials used. The
Report also includes guidance on permissible tolerances in
pultruded plate geometry.

10.5 STORAGE OF MATERIALS


Thick fibre composite plates and NSM reinforcement are
usually delivered to site in the lengths required for installation. Thinner plates and fabrics are delivered in the form of
a long roll, which can be cut easily to the required lengths at
site. Materials should be stored at site in such a way that
damage or contamination is avoided.
Adhesives should be stored in dry conditions in accordance
with the manufacturers instructions, paying particular
attention to the specified maximum and minimum storage
temperatures. Adhesive and material delivery dates should
be recorded and these items should be used in rotation.

10.6 SITE CONDITIONS

Figure 40: Pull-off specimen after removal from concrete


surface.

10.3.2 Slots in concrete surface for NSM material


Before starting to form slots for NSM reinforcement, a
covermeter survey should be carried out to check the position and depth of the existing steel reinforcement in the area
to be strengthened, to avoid damaging it. Slots are formed by
making parallel cuts in the surface to the required depth, at
an appropriate distance apart, and removing the intervening
concrete using a chisel or similar. The slots should be cleaned,
using a vacuum cleaner or high-pressure air, to remove any
loose material but otherwise should not require any further
preparation to ensure that the adhesive has adequate bond.

Temperature, relative humidity and surface moisture at the


time of installation can affect the performance of the FRP
system. It is therefore necessary to maintain the appropriate
environment in the work area during surface preparation,
application of the adhesive and the subsequent curing period.
Environmental control during surface preparation generally
consists of a system to extract dust from the work area and
the exclusion of any material that might contaminate the
prepared surface. A clear access path should be maintained
from the area where the adhesive is applied on the plates to
the location of the concrete surface to which the plates are to
be applied. This is to minimise the risk of contamination of
the adhesive surface while the plate is being handled.
During the curing period it is necessary to maintain the temperature in the adhesive at an appropriate value and within
specified limits. Exceeding the maximum specified temperature may result in a joint with poor long-term properties.
Curing temperatures below the specified minimum may result
in an adhesive with a low strength. Of equal importance is
keeping the work dry.

73

Design guidance for strengthening concrete structures using fibre composite materials

10.7 MIXING AND APPLICATION OF


ADHESIVE
10.7.1 General
All equipment used for the mixing and application of the
adhesive and materials should be kept clean and maintained
in good operating condition. All operatives should be suitably
trained in the use of such equipment.

10.7.2 Application to substrate prior to plate installation


Where the concrete surface is to be strengthened using FRP
plates, the mixed adhesive is applied to the bonding area by
hand, using plastering techniques (see Figure 42). The thickness of the adhesive should be maintained at 12mm.

The mixing and application of the adhesive should be strictly


in accordance with the manufacturers instructions. (Figure
41 shows adhesive being mixed using a power tool with a
suitable attachment.) In particular, the amounts of materials
mixed at any one time should not exceed the specified
amounts, as larger volumes will lead to higher temperatures
being generated, which will reduce the pot life. Resin and
hardener have to be mixed together in defined proportions or
the properties of the cured adhesive will be impaired. Hence
prebatched quantities of resins and hardeners should be used.
Figure 42: Application of adhesive to concrete surface.

10.7.3 Application to FRP plates


Before installation, FRP plates should be checked visually
for signs of damage, such as cracks or delamination. The
surface of the plate should be prepared immediately before
application of the adhesive, in accordance with the manufacturers recommendations. This may involve light abrasion
and cleaning with a solvent. Some materials are manufactured with an additional peel ply, which, on removal,
exposes a clean surface with the appropriate roughness. This
is the preferred approach since no additional treatment at site
is required.
The adhesive layer should be applied to the plates to form a
slightly convex profile across the plate. The extra thickness
along the centre-line helps to reduce the risk of void
formation. A method of application is shown in Figure 43.

Figure 41: Mixing adhesive.

The components should be thoroughly mixed together. Some


adhesives are supplied with resin and hardener of different
colours. This makes it easier to check that thorough mixing
has been achieved.
The volume of adhesive mixed at one time must be such that
it may be applied and the surfaces brought together within
the pot life of the adhesive. Any adhesive remaining at the
end of the specified pot life must be discarded.

74

Figure 43: Application of adhesive layer on to fibre


composite plate.

Workmanship and installation

10.7.4 Application to substrate prior to fabric installation


Where the concrete surface is to be strengthened using FRP
fabric, the bonding adhesive is applied using a hand-held
foam roller or brush. This should be evenly applied to
saturate the concrete surface and promote adhesion of the
fabric material.
10.7.5 Application to FRP fabrics
Fabric can be readily cut to size using simple tools (see
Figure 44). Dry fabric can be directly applied to the resinsaturated concrete surface without adhesive being applied to
the fabric. For wet fabric, the resin is applied to the fabric
before it is installed. This resin can be applied to the fabric
using hand-held foam rollers (see Figure 45), brushes or impregnation machines (see Figure 46). Alternatively, vacuumassisted resin infusion can be used to form the composite in
situ. The fibres are applied to the structure dry. The area is
sealed with a rubber sheet and a vacuum used to draw in the
resin.

Figure 46: Impregnation of fabric.

10.7.6 Inserting adhesive into slots for NSM


reinforcement
When NSM reinforcement is installed on the top surface on
a member, adhesive is simply poured into the slot to a depth
equal to approximately the eventual mid-section of the NSM
rod or strip. The adhesive needs to be sufficiently fluid to
flow into the slot without entrapping air. For installation overhead or on vertical surfaces, a stiffer adhesive is required
which will not slump significantly. Installation will be by
means of an adhesive gun.

10.8 ASSEMBLY AND VISUAL


INSPECTION
10.8.1 Installation of FRP plates
Figure 44: Cutting fabric.

Immediately after application of the adhesive the fibre


composite plate should be brought into contact with the
concrete substrate. There is sufficient grab in the adhesive
to hold the fibre composite material in position, and no other
temporary support is usually needed.
Even pressure is applied by roller (as shown in Figure 47)
starting at one end along the longitudinal centre-line and
working outwards to expel excess adhesive at the edges and
to produce an even glue line. A final adhesive thickness of
1.52mm is ideal in most cases. Excess adhesive is removed
using scrapers, cloths and solvents.

Figure 45: Applying resin using roller.

75

Design guidance for strengthening concrete structures using fibre composite materials

If required, a second and further layers of fabric can be applied


in a similar fashion, making sure that each successive layer
is fully saturated with resin. Finally, a layer of epoxy adhesive may be applied to encapsulate and protect the composite
material. Alternatively, the fabric can be impregnated with
resin, i.e. wet fabric, and then wrapped around the member.
As before, the surface should be rolled to remove wrinkles
and to expel air.

Figure 47: Installing FRP plates, using a roller to apply


pressure.

Where plates are lapped, the minimum overlap, in the


direction of the fibres, should be in accordance with the
manufacturers recommendations, but not less than 200mm.
Chapter 6 gives more detailed information on lap lengths.
The spacing of FRP plates on the soffit or top surface of a
slab should not exceed 0.2 x span or 5 x slab thickness.
Immediately after assembly, the joint should be inspected.
The aim is to check that a continuous and uniform layer of
adhesive is visible. In some situations the soundness of the
installed adhesive layer can be checked by tapping the
composite with a small object such as the edge of a coin.
Voids or gaps give a characteristic sound. Further information is given in Section 4.3 of TR57(5). If defects are found
techniques such as vacuum filling with a suitable resin or
plate overlapping could be used as a repair. Further information is given in Section 6.2 of TR57.
10.8.2 Installation of FRP fabrics

Figure 49: Wrapping fabric round column.

The dry fabric is wrapped tightly over the concrete substrate,


avoiding any wrinkles. Figure 48 shows fabric being wrapped
round an arched member and Figure 49 shows the wrapping
of a column. After application, the fabric is rolled to force
the adhesive through the fibres and to expel any air (see
Figure 50).

Figure 50: Rolling fabric to consolidate layers.

Figure 48: Wrapping fabric round an arched member.

76

The minimum overlap for fabric materials, in the direction of


the fibres, should be in accordance with the manufacturers
recommendations, but should not be less than 200mm. The
FIB guide(90) suggests a maximum of five layers in a given
direction. However, some suppliers suggest that more layers
may be used. Caltrans (the California Department of Transportation) permits up to 14 layers. Advice should be sought
from the supplier.

Workmanship and installation

10.8.3 Installation of NSM reinforcement


The NSM reinforcement is placed in the slot and pressed into
the adhesive, which partly fills the slot, to the correct depth.
The remaining void in the slot is then filled with more
adhesive (taking care to avoid trapping air) and finished
flush with the concrete surface.

10.9 CONTROL SAMPLES


Tests should be carried out on samples obtained from each
batch of adhesive and on the composite materials for testing
by an independent laboratory to confirm the properties of the
materials used. These tests should be in accordance with
agreed national or international standards, as detailed in
Appendix B of TR57(5).
Figure 51 shows a double-lap shear test, which is closely
related to the actual site use of fibre composite plates. As the
embedded studs may cause premature failure of the concrete
blocks, alternative methods are being developed for loading
the specimens, for example by the University of Glamorgan(177). The test can be carried out using a standard universal
testing machine and can be carried out, relatively easily, at
different temperatures. Compatibility of fibre, or composite,
and adhesive can be tested, as well as adhesion to concrete.
Additionally, pull-off tests can be performed to check the
adhesion of the adhesive to the substrate.

can guarantee the soundness of an application. Full details of


available tests are given in TR57(5). The most common is
acoustic sounding (hammer tapping). Thermography may be
used to survey large areas. Other methods, such as ultrasonic
testing, are being developed. However, there are currently no
non-destructive methods that are capable of detecting poor
adhesion, which might lead to failure of the joint in the long
term. To provide assurance about long-term performance,
additional pull-off dollies could be installed at the time of
strengthening. Pull-off tests could be performed on these at
various times in the future to monitor the adhesion between
the adhesive and the substrate. Similarly, additional doublelap shear test specimens could be prepared at the time of
strengthening, placed securely on site, and tested at various
times in the future to monitor the adhesion between the FRP
material and the adhesive.
Some documents, such as TR57(5) and ACI 440.2(2), suggest
the extent of delamination that may be acceptable. However,
the acceptable extent will be very dependent on the type of
strengthening and the location in the structure. For example,
an area of delamination in the wrapping of a column will
probably have a limited effect on the performance while
delamination of a plate on the soffit of a beam, particularly
at points of high shear, will have a significant effect.
Currently there are no techniques for inspecting NSM strengthening systems after installation, apart from a visual check for
major voids.
For major structures, it may be appropriate to install instrumentation prior to the strengthening. Measurements of the
difference in the response of the structure under a load test
before and after strengthening can be compared with predictions and the instrumentation used to monitor changes with
time.

10.11 APPLICATION OF OVER-COATINGS


When over-coatings are to be applied, these should be compatible with the underlying composite material and approved
for use by the manufacturer. These over-coatings may be
applied for the following reasons:

Figure 51: Double-lap shear test.

1504(17)

BS EN
will be a product standard for materials for
the repair and protection of concrete. (Currently some parts
have yet to be published as British Standards.) It will detail
the required properties of the materials and the tests that are
required to demonstrate conformity.

10.10 NON-DESTRUCTIVE TESTS


Various non-destructive tests may be used to inspect a
completed and cured bond for surface-bonded materials.
However, there is no regime of non-destructive testing that

Fire protection. Regulations may require the application


of an over-coat layer, which has been tested on the fullycured composite system.
Protection against vandalism. Where the FRP material
may be vulnerable to damage, it may be encapsulated in
a cementitious or epoxy mortar either spray- or handapplied.
Appearance. A cosmetic over-coating can be applied to
the composite material to match the existing structure.
Protection against ultraviolet radiation. The manufacturer should be consulted for advice on the UV resistance
of the FRP. If the manufacturer recommends UV protection, a cementitious, or other, over-coating can be applied.

77

Design guidance for strengthening concrete structures using fibre composite materials

Reducing solar gain. In exposed situations the FRP


could be painted white or shrouded to reduce solar gain
and the consequent temperature rise.
Other reasons. In certain circumstances the FRP material may be encapsulated by other structural finishes or,
on bridge decks, covered by the waterproofing.

Figure 52 shows a sprayed mortar over-coat being applied to


FRP plates on the soffit of Dudley Port Bridge.

Figure 52: Spray application of mortar over-coating.

10.12 IDENTIFICATION/WARNING SIGNS


There may be a risk that the fibre composite will be damaged
by other work carried out on the structure. For example,
holes drilled though to take the fixings for a false ceiling
would seriously affect the capacity of the fibre composite;
because of stress-concentrations around the hole, drilling
will lead to a loss in strength equivalent to a reduction in
section of two- or three-hole diameters. When FRP is bonded
to the upper surface of a member and covered by a surfacing
with a limited life, removal of the surfacing may lead to
damage of the FRP. In such cases suitable identification/warning plates or other markings should be fixed on
or adjacent to the composite (see Figures 53 and 54). Where
an over-coating layer is applied, the plates should, where
possible, be placed on the exposed surface.

Figure 54: Examples of proposed warning plates fixed to


structure adjacent to strengthened area.

10.13 RECORDS
Detailed records should be kept of the work carried out, as
required under the CDM Regulations(26), and added to the
Health and Safety File, which should include details of any
future inspection and testing regime that is considered appropriate. Details of the records that should be kept, along with
suggested proformas and check lists, are given in TR57(5).
Figure 53: Example of warning printed on carbon fibre plate.

78

11 LONG-TERM INSPECTION,
MONITORING AND MAINTENANCE

11.1 INSPECTION AND MONITORING


REGIME
Full details on the requirements for inspecting and monitoring structures strengthened with FRP are given in TR57(5),
which includes checklists for the aspects to be considered at
various stages and a standard proforma for recording inspection data. The following Sections summarise the information
in TR57, which should be read in parallel with this document.
As with all structural elements there will be a need to check
the fibre composite strengthening system as part of the regular
inspection and monitoring of the structure. Such inspections
are already carried out for bridges, with general (visual)
inspections annually and detailed inspections every six years
or so. However, buildings are rarely inspected on a regular
basis, inspections often being carried out only when there is
a change of use or of ownership. It is strongly recommended
that all building owners should instigate a regular inspection
regime for strengthened elements.
Information on the materials used in the strengthening should
be included in the Health and Safety File for the structure.
This File should also include details of any initial faults in
the fibre composite strengthening, such as minor areas of
delamination, and should indicate those regions of the
strength-ening that are critical, such as anchorage zones. The
structural engineer responsible for designing the strengthening should indicate the action to be taken in the event of
any likely forms of damage to the composite material. An
example of this might be damage to fibre composite material
on the soffit of a bridge following impact by an over-height
vehicle. The action to be taken will be specific to the particular structure as it will depend on the amount of damage
and the extent to which the structure has been strengthened.
Hence no general guidance can be given in this Report.
It is strongly recommended that additional samples of the
fibre composite material should be bonded to the structure
away from the region to be strengthened. (This approach has
been adopted on a number of structures including the Barnes
Bridge in Manchester and the John Hart Bridge in British
Columbia (see Section 4.3).) Additionally, or alternatively,
FRP can be bonded to concrete samples, such as short
beams, which can be stored on or adjacent to the structure.
Samples can be inspected and tested as part of the inspection
regime. To aid inspection, some or all of the samples should
not be covered with any protective layer. They should thus

indicate a lower bound to the performance of the composites


bonded to the main structure. Details should be included in
the Health and Safety File along with recommendations for
the frequency of testing.
Finally, the Health and Safety File should include details of
any instrumentation that was installed as part of the strengthening exercise, along with any data obtained before and after
strengthening.

11.2 FREQUENCY OF INSPECTIONS


The intervals between inspections recommended below,
taken from Chapter 1 of TR57(5), should be taken only as a
guide. Structures in aggressive environments will require
more frequent inspection. Special structures may require a
special inspection regime, the frequency and extent of which
being determined by a risk assessment.
Routine, visual inspection
The recommended intervals for routine visual inspection are
as follows:
Bridges
Buildings
Other structures

Every year
Every year
Depends on the use of the structure
but ideally every year.

Detailed inspection with testing


In the absence of other guidance, detailed inspections should
be carried out at intervals as follows:
Bridges
Buildings

Other structures

At least every six years


At change of occupancy or change
of use, when structural work or
refurbishment is carried out on the
building, but at intervals not exceeding ten years
Depending on the use of the structure but at least every ten years.

Detailed inspections should be carried out more frequently


in the first few years after installation, to give the owner of
the structure confidence that the strengthening has been
carried out satisfactorily.

79

Design guidance for strengthening concrete structures using fibre composite materials

11.3 ROUTINE VISUAL INSPECTION


Information on routine visual inspection is given in Section
5.2 of TR57. The surface of the fibre composite should be
inspected for signs of crazing, cracking or delamination,
which would indicate some level of overall deterioration.
The composite should be inspected for local damage, for
example caused by impact or abrasion. In addition, of course,
the inspection should look for signs of the deterioration of
the concrete structure itself, such as additional cracking or
corrosion.
Where the composite has been covered with over-coating, it
will not be possible to directly inspect the composite.
Damage to the protective layer will suggest the possibility of
damage of the composite. In general, it will not be appropriate to remove the protective layer as this may cause damage
to the fibre composite. Thus any inspection of the composite
will have to be limited to the control samples.
Identification/warning labels (see Section 10.12 and Figures
54 and 55) should be checked and missing ones should be
replaced. This is particularly important where there is the
likelihood of future work that could damage the fibre composite material, such as the installation of fixings for services.
The fibre composite may have been covered by paint or
other form of protective layer, e.g. for protection from ultraviolet light, which will have a limited life. It will be necessary
to check the condition of this layer and to replace it when
required, in accordance with the suppliers recommendations,
with a material that is compatible with the fibre composite.

11.4 DETAILED INSPECTION


Information on detailed inspection and testing is given in
Section 5.3 of TR57. Debonding of the fibre composite
material from the concrete may be determined by tapping or
thermography, as indicated in Section 10.10. However, there
are currently no simple, non-destructive tests that can be
used to assess the condition of the adhesive. This is best
determined by carrying out pull-off tests on the control
specimens at regular intervals. These tests should be carried
out as part of the detailed inspection, though there may be a
requirement to test samples more frequently, at least during
the early period after the strengthening.

80

Instrumentation may have been installed as part of the


assessment process, for example to measure strains due to
live loading on the structure. In addition, instrumentation
may have been installed on the structure at the time of
strengthening, to enable the response to be compared with
that predicted. Such instrumentation can be used to indicate
changes in the response. If significant changes are observed,
it will be necessary to identify whether they are due to
changes in the strengthening system (such as delamination)
or due to overall changes in the concrete structure (such as
additional cracking or corrosion) so that appropriate action
can be taken. It will be necessary for the structure to be reanalysed by a structural engineer to determine what remedial
action may be required.
When local areas of damaged composite are identified, they
may be repaired by techniques such as vacuum filling with a
suitable resin (taking care not to further damage the material) or plate overlapping. When major damage is identified,
such as peeling and debonding of large areas, it may be
necessary to remove the defective material and adhesive.
The defective material should be removed over a sufficiently
large area such that material on the periphery is fully bonded.
The concrete surface should then be prepared again and
further FRP installed. It will be necessary to provide an
adequate overlap between the new and old material at the
periphery of the repaired area. Where these repair techniques
are used, it is crucial to check the compatibility of the repair
material with the materials already in place. In addition to
compatibility, the repair material must have similar characteristics to the material in place. Such characteristics include
fibre orientation, volume fraction, strength, stiffness and overall thickness. Some additional information on repair is given
in Section 6.2 of TR57.

11.5 MAINTENANCE
The nature of FRP materials means that they should need
little or no maintenance while in service. However, as indicated in Section 2.3 of TR57, moisture is one of the most
damaging elements and so all gutters, drains, etc., must be
kept clear of debris, so that rainwater is carried off the
structure and away from the FRP. If any cleaning is carried
out near the FRP, it must be checked that any solvents used
will not cause damage. Cleaning techniques such as waterjetting or grit-blasting are not appropriate as they are likely
to cause damage to the FRP.

REFERENCES

1.

2.

3.

4.

5.

6.

7.

8.

9.

10.

CONCRETE SOCIETY. Design guidance for


strengthening concrete structures using fibre composite materials, Technical Report 55, The Concrete
Society, Camberley, 2000, 71pp.
AMERICAN CONCRETE INSTITUTE. ACI 440.2R,
Guide for the design and construction of externally
bonded FRP systems for strengthening concrete
structures, American Concrete Institute, Farmington
Hills, Michigan, USA, 2002.
ISIS CANADA. Strengthening reinforced concrete
structures with externally-bonded fibre-reinforced
polymers, Manual No. 4, Report ISIS-MO5-00, ISIS
Canada, University of Manitoba, Winnipeg, Canada,
2001.
CANADIAN STANDARDS ASSOCIATION. CSAS806-2, Design and construction of building components with fibre-reinforced polymers, Mississauga,
Ontario, 2002.
CONCRETE SOCIETY. Strengthening concrete
structures using fibre composite materials: acceptance,
inspection and monitoring, Technical Report 57, The
Concrete Society, Camberley, 2003, 48pp.
CONSTRUCTION INDUSTRY RESEARCH AND
INFORMATION ASSOCIATION. Strengthening
metallic structures using externally-bonded fibrereinforced polymers, Report C595, CIRIA, London,
2004, 234pp.
CONCRETE SOCIETY. Diagnosis of deterioration in
concrete structures, Technical Report 54, The Concrete
Society, Crowthorne, 2000, 72pp.
INSTITUTION OF STRUCTURAL ENGINEERS.
Appraisal of existing structures (Second Edition), The
Institution of Structural Engineers, London, 1996,
106pp.
HIGHWAYS AGENCY. BA 30/94, Strengthening of
concrete highway bridges using externally bonded
plates, Design manual for roads and bridges. Volume
3, Highway structures: inspection and maintenance.
Section 3, Repair. Part 1, Department of Transport,
London, 1994.
HIGHWAYS AGENCY. BD 84/02, Strengthening of
concrete bridge supports for vehicle impact using
fibre reinforced polymers, Design manual for roads
and bridges. Volume 1, Highway structures: approval
procedures and general design. Section 3, General
design. Part 16, Department of Transport, London,
2002.

11.

12.
13.
14.

15.

16.

17.

18.

CANNING, L. and LUKE, S. Field testing and long


term monitoring of West Mill Bridge, In: Hollaway,
L.C., Chryssanthopoulos, M.K. and Moy, S.S.J. (Eds),
Advanced polymer composites for structural applications in construction, Woodhead Publishing Limited,
2004, pp.683692.
ANON. Packing more road into parkway, Engineering
News Record, 12 May 1997, pp.3031.
KLIGER, H. Repair of parking structures, FRP International, Vol. IV, Issue 4, Autumn 1996, pp.34.
ALEXANDER, J.G.S. and CHENG, J.J.R. Field application and studies of using CFRP sheets to strengthen
concrete bridge girders. In: El-Badry, M. (Ed.),
Advanced composite materials in bridges and structures, Canadian Society for Civil Engineering,
Montreal, 1996, pp.465472.
LUKE, S., GODMAN, S. and HAMNEVOLL, H.
Strengthening the Qafco Prill Tower, Qatar, with FRP
plates, CONCRETE, February 2002, pp.3233.
HOOKS, J.M. and COOPER, J.D. Applications of
FRP composites for bridge rehabilitation and strengthening in the USA, In: Forde, M.C. (Ed.), Proceedings
of Tenth International Conference on Structural
Faults and Repair 2003, Technics Press, Edinburgh,
2003, available as CD.
BRITISH STANDARDS INSTITUTION. BS EN
1504. Products and systems for the protection and
repair of concrete structure. Definitions, requirements, quality control and evaluation of conformity;
Part 1, Definitions; Part 2, Surface protection; Part 3,
Structural and non-structural repair; Part 4, Structural
bonding; Part 5, Concrete injection; Part 6, Grouting
to anchor reinforcement or to fill external voids; Part
7, Reinforcement corrosion prevention; Part 8,
Quality control and evaluation of conformity; Part 9,
General principles for use of products and systems;
Part 10, Site application of products and systems and
quality control of the works, BSI, London. (Note that
only Parts 1 and 10 are currently available as BS ENs.
The remaining parts are at various stages of
development.)
BALAGURU, P. and TOUTANJI, H. Recent advances
in inorganic polymer composites for repair and
rehabilitation, In: Grantham, M., Rendell, F. and
Jauberthie, R., (Eds), Proceedings of Concrete Solutions, First International Conference on Concrete
Repair, St. Malo, France, 1517 July 2003.

81

Design guidance for strengthening concrete structures using fibre composite materials

19.

20.

21.
22.

23.

24.

25.

26.
27.
28.
29.

30.

31.

32.

33.

34.

82

HULL, D. and CLYNE, T.W. An introduction to composite materials, Cambridge University Press, 1996,
326pp., ISBN 0-521-38855-4.
MATTHEWS, F.L. and RAWLINGS, R.D. Composite
materials: engineering and science, Chapman & Hall,
1994, 470pp.
MEIER, H. CFRP L-shaped plates, Schweitzer Ingenieur und Architekt, No. 43, 1998.
CZADERSKI, C. Shear strengthening with CFRP Lshaped plates, Schweitzer Ingenieur und Architekt,
No. 43, 1998.
MAYS, G.C. and HUTCHINSON, A.R. Adhesives in
civil engineering, Cambridge, Cambridge University
Press, 1992.
INSTITUTION OF STRUCTURAL ENGINEERS. A
guide to the structural use of adhesives, The Institution of Structural Engineers, London, 1999, 51pp.
LONDON UNDERGROUND LIMITED. Fire safety
performance of materials used in the Underground,
Engineering Standard E 1042 A3, London, March
1998, 12pp.
The Construction (Design and Management) Regulations, SI 1994/3247, HMSO.
Control of Substances Hazardous to Health Regulations, SI 437, 1999. The Stationery Office.
PARKER, D. Sticking to the task, New Civil Engineer,
11 July 1996, p.22.
HOLLAWAY, L. and LEEMING, M.B. (Eds). Strengthening of reinforced concrete structures using externally
bonded FRP composites in structural and civil
engineering, Woodhead Publishing, 1999, pp.327.
GOLD, C.A. and MARTIN, A.J. Refurbishment of
concrete buildings: structural and services options,
Guidance Note GN 8/99, BSRIA, Bracknell, 1999,
114pp.
LUKE, S. Strengthening of structures using carbon
fibre reinforced polymer plates, In: Proceedings of
Reinforced Plastics 98 Conference, Singapore, 1998.
TAYLOR, M.J., LUKE, P.S., COLLINS, S. and
DARBY, J.J. Using advanced composites in strengthening and rehabilitation of rail bridges, In: Forde,
M.C. (Ed.), Proceedings of Second International
Conference on Railway Engineering, London, May
1999. ECS Publications, Edinburgh.
GARDEN, H.N. Carbon fibre composites as structural
reinforcement in a nuclear installation. In: Virdi, K.S.,
Matthews, R.S., Clarke, J.L. and Garas, F.K. Abnormal
loading on structures, E & FN Spon, London, 2000,
pp.297305.
LUKE, S. Using composites in the strengthening and
rehabilitation of infrastructure assets, In: Proceedings
of Conference on Rehabilitation of Piping and
Infrastructure, Newcastle, March 1999.

35.

36.

37.

38.

39.
40.
41.
42.
43.

44.
45.
46.

47.
48.

49.

50.

51.

52.

RUSSELL, P. and LOMAX, K. Wrapping the Harbour


City car park and other projects, In: Forde, M.C. (Ed.),
Proceedings of Tenth International Conference on
Structural Faults and Repair 2003, Engineering
Technics Press, Edinburgh, 2003, available as CD.
RUSSELL, P and MODI, A. Wrapping the Harbour
City car park, Manchester, Concrete, Vol. 35, No. 6,
June 2001, p.24.
FARMER, N. Near-surface-mounted carbon fibre
reinforcement, Concrete, Vol. 37, No. 1, January 2003,
pp.2021.
FARMER, N. Near-surface-mounted reinforcement
preserves car park for Merseyside Police, Concrete,
Vol. 37, No. 9, October 2003, pp.3638.
SOUDAIN, M. Cost cuts on a plate, New Civil
Engineer, 12 March 1998, p.12.
ANON. Passing the plate, New Civil Engineer, 13
February 1997, p.9.
ANON. Bridge strengthening, The Structural
Engineer, 19 May 1998, pp.A11,12.
SMITH, K. Mouchel hands out repairs on a plate,
Construction News, 16 October 1997.
LANE, J.S., LUKE, S. and LEEMING, M.B. Strength
on a plate, Surveyor, Vol. 185, No. 5476, 19 March
1998.
PARKER, D. Keeping damage under wraps, New Civil
Engineer, 20/27 August 1998, pp.3132.
ANON. Bridge strengthening, Surveyor, 7 January
1999, p.23.
FARMER, N.S. Glade Bridge repair: carbon fibre
composites case proven, In: Forde, M.C. (Ed.), Proceedings of Second International Conference on
Railway Engineering, London, May 1999. ECS
Publications, Edinburgh.
ANON. Sika quadruples strength, Concrete, Vol. 33,
No. 1, January 1999, p.41.
BARTON, R. Upgrading bridge with carbon-fibre
composites, The Structural Engineer, Vol. 77, No. 5, 2
March 1999, p.30.
SADKA, B. Strengthening bridges with fibrereinforced polymers, Concrete, Vol. 34, No. 2,
February 2000, pp.4243.
DENTON, S.R., FARHADI, B., BENT, T. and
SHAVE, J.D. Bridge column strengthening using FRP
a case study, In: Proceedings of International
Conference Structural Faults and Repair, London,
2001, Engineering Technics Press, Edinburgh, 2001.
KENDALL, D. Reinforcement of concrete columns
using preformed FRP shells, In: Proceedings of
International Conference Structural Faults and
Repair, London, 2001, Engineering Technics Press,
Edinburgh, 2001.
PINZELLI, R. FRP shells for strengthening bridge
columns, Concrete, Vol.35, No. 6, June 2001,
pp.2223.

References

53.

54.
55.

56.

57.

58.

59.

60.
61.

62.

63.

64.

LUKE, S. and CANNING, L. Strengthening highway


and railway bridge structures with FRP composites
Case studies, In: Hollaway, L.C., Chryssanthopoulos,
M.K. and Moy, S.S.J. (Eds), Advanced polymer
composites for structural applications in construction,
Woodhead Publishing Limited, 2004, pp.747754.
DENTON, S.R. Mastering the mechanics, Bridge
Design and Engineering, 2nd Quarter 2002, pp.4849.
DREWETT, J. A92 Tay Road Bridge: wrapping north
approach viaduct columns with FRP, Concrete,
Volume 38, Number 5, May 2004, pp.17 & 19.
RICHARDSON, M. The development of composite
materials for strengthening structures, In: Hollaway,
L.C., Chryssanthopoulos, M.K. and Moy, S.S.J. (Eds),
Advanced polymer composites for structural applications in construction, Woodhead Publishing Limited,
2004, pp.755762.
CHACOS, G.P. Parking structure modified to increase
headroom, In: Grantham, M., Rendell, F. and Jauberthie, R., (Eds), Proceedings of Concrete Solutions,
First International Conference on Concrete Repair,
St-Malo, France, 1517 July 2003.
IGNOUL, S., BROSENS, K. and VAN GEMERT, D.
Strengthening of concrete structures with externally
bonded reinforcement: design concepts and case
studies, In: Grantham, M., Rendell, F. and Jauberthie,
R., (Eds), Proceedings of Concrete Solutions, First
International Conference on Concrete Repair, StMalo, France, 1517 July 2003.
STEINER, W. Strengthening of structures with FRP
strips, In: El-Badry, M.M. (Ed.), Advanced composite
materials in bridges and structures, Canadian Society
for Civil Engineering, Montreal, 1996, pp.407417.
BLASZAK, G. Strengthening of airport parking garage,
FRP International, Vol. VI, Issue 4, Autumn 1998, p.5.
DORTZBACH, J. Carbon fiber reinforcing polymers
as negative moment reinforcing in repair of composite
steel parking deck, In: Dolan, C.W., Rizkalla, S. and
Nanni, A. (Eds), Fibre reinforced polymer reinforcement for reinforced concrete structures, SP-188,
American Concrete Institute, Farmington Hills,
Michigan, USA, 1999, pp.417428.
LILLISTONE, D. and JOLLY, C.K. Concrete-filled
fibre reinforced plastic circular columns, In: Composite Construction Conventional and Innovative,
IABSE Report, Zurich, 1997, pp.759764.
ANON. Reinforcement of structures with carbon
fibres. Freyssinet Magazine, December 1996/January
1997.
HUTCHINSON, R. Fibre for health; strengthening the
Maryland Street Bridge, Innovator Newsletter of
ISIS Canada, February 2000.

65.

66.
67.

68.

69.

70.
71.

72.

73.

74.

75.

76.

77.
78.

79.

MEIER, U., DEURING, M., MEIER, H. and


SCHWEGLER, G. Strengthening of structures with
advanced composites. In: Clarke, J.L. (Ed.), Alternative materials for the reinforcement and prestressing
of concrete. Blackie Academic and Professional,
Glasgow, 1993, pp.151171.
NANNI, A. CFRP strengthening, Concrete International, Vol. 19, No. 6, June 1997, pp.1923.
MEIER, U. Repair using advanced composites,
Composite Construction Conventional and Innovative, IABSE Report, Zurich, 1997, pp.113124.
LYNCH, M.J. and DUCKETT, W.G. A30 Bible
Christian Bridge, Cornwall Advanced composite
material column wrapping trial, Highways Agency
Maintenance Conference, Nottingham University, 8
September 1998.
ANON. Rehabilitation following earthquake disaster.
FRP International, Vol. III, Issue 4, Autumn 1995,
pp.45.
ANON. Caltrans now permits composite wraps,
Engineering News Record, 25 December 1995.
CERCONE, L. Highway bridge pier column emergency repair, Advanced Materials & Composites News,
Vol. 22, No. 5, March 2000, pp.56.
NEALE, K. Rehabilitation of columns of a highway
overpass using fiber composite materials, FRP International, Vol. IV, Issue 4, Autumn 1996, p.4.
NEALE, K.W. and LABOSSIRE, P. Fiber composite
sheets in cold climate rehab, Concrete International,
Vol. 20, No. 6, June 1998, pp.2224.
NEALE, K.W. and LABOSSIRE, P. State-of-the-art
report on retrofitting and strengthening by continuous
fibre in Canada, In: Non-metallic (FRP) reinforcement
for concrete structures, Volume 1, Japan Concrete
Institute, 1997, pp.2539.
SIWOWSKI, T. FRP composites or external prestressing in bridge strengthening comparison, In:
Forde, M.C. (Ed.), Proceedings of Tenth International
Conference on Structural Faults and Repair 2003,
Engineering Technics Press, Edinburgh, 2003, available as CD.
ROBERTS, J.E. Composite construction in California
bridge seismic retrofitting, Composite Construction
Conventional and Innovative, IABSE Report, Zurich,
1997, pp.943952.
ANON. Repairing columns without using forms, Concrete International, Vol. 21, No. 3, March 1999, p.82.
KARBHARI, V.M. and SEIBLE, F. Fiber-reinforced
polymer composites for civil infrastructure in the
USA, Structural Engineering International, Vol. 9,
No. 4, November 1999, pp.274277.
OKAMOTO, T. Aramid tape for seismic strengthening, FRP International, Vol. IV, Issue 3, Summer
1996, p.3.

83

Design guidance for strengthening concrete structures using fibre composite materials

80.
81.

82.

83.

84.

85.
86.
87.

88.

89.

90.

91.

92.

93.

94.

95.

84

DIAL, S. GFRP for smoke stack, FRP International,


Vol. VI, Issue 2, Spring 1998, pp.34.
FUKUYAMA, H., NAKAI, H., TANIGAKI, M. and
UOMOTO, T. JCI state-of-the-art on retrofitting by
CFRM, Part 1, Materials, construction and application, Proceedings of the Third Symposium on Nonmetallic (FRP) Reinforcement for Concrete, Japan
Concrete Institute, 1997, Vol. 1, pp.605612.
NEALE, K.W. FRPs restore 1910 water chamber to
original capacity, Innovator Newsletter of ISIS
Canada, February 1999.
DEANE, C. Deep water jetty for Humber Sea
Terminal, Concrete Engineering International, Vol. 8,
No. 1, Spring 2004.
PHAIR, M. Navy looks to composites for low-cost
pier upgrade, Engineering News Record, 12 January
1998, p.17.
GEE, D. Underwater applications of FRP, FRP
International, Vol. VI, Issue 4, Autumn 1998, p.2.
ANON. Structural reinforcement from MBT FEB,
Concrete, Vol. 35, No. 1, January 2001, pp.1213.
EL-HACHA, R., EL-BADRY, M. and ABDALLA, H.
Strengthening of prestressed concrete railway ties using
composite straps, In: El-Badry, M.M. (Ed.), Advanced
composite materials in bridges and structures, Canadian Society for Civil Engineering, Montreal, 1996,
pp.489496.
BRITISH STANDARDS INSTITUTION. BS 8110,
Structural use of concrete, Part 1: 1997, Code of
practice for design and construction, Part 2: 1985, Code
of practice for special circumstances, BSI, London.
BRITISH STANDARDS INSTITUTION. ENV 19921, Eurocode 2: Design of concrete structures, Part 1:
General rules and rules for buildings, BSI, London,
1992.
FEDERATION INTERNATIONAL DU BETON.
Externally bonded FRP reinforcement for RC
structures, Bulletin 14, FIB, Lausanne, 2001, 138pp.
JAPAN SOCIETY OF CIVIL ENGINEERS. Recommendations for upgrading concrete structures with
use of CFRP, Japan Society of Civil Engineers, 2001.
TLJSTEN, B. Strengthening of existing concrete
structures: Design guidelines, Second Edition,
Division of Structural Engineering, Lule University
of Technology, Sweden, 2003.
CONSTRUCTION INDUSTRY RESEARCH AND
INFORMATION ASSOCIATION. Fibre-reinforced
polymer composites in construction, CIRIA, London,
2002.
CLARKE, J.L. (Ed.). Structural design of polymer
composites EUROCOMP design code and handbook, E & F N Spon, London, 1996, 751pp.
BRITISH STANDARDS INSTITUTION. BS 5400,
Steel, concrete and composite bridges, Part 4: 1990,
Code of practice for design of concrete bridges, BSI,
London.

96.

97.

98.

99.

100.

101.

102.

103.

104.

105.

106.

107.

HIGHWAYS AGENCY. BD 44/95. The assessment of


concrete highway bridges and structures, Design
manual for roads and bridges. Volume 3, Highway
structures: inspection and maintenance. Section 4,
Assessment. Part 14, Department of Transport, London,
1995.
HIGHWAYS AGENCY. BD 37/01, Loads for highway bridges, Design manual for roads and bridges.
Volume 1, Highway structures: approval procedures
and general design. Section 3, General design. Part 14,
Department of Transport, London, 2001.
STECKEL, G.L. Environmental durability of Fyfe
Company SHE 51/Tyfo E E-glass reinforced epoxy
composite, Aerospace Report No. ATR-99(7455)-3,
The Aerospace Corporation, USA, 1999.
BRITISH STANDARDS INSTITUTION. BS 476:
Fire tests on building materials and structures, Part
20: Method for determination of the fire resistance of
elements of construction (general principles), 1987,
BSI, London.
TOURNEUR, C. First application of FRP sheet in
France, FRP International, Vol. VI, Issue 2, Spring,
1998, pp.34.
PRIESTLEY, M.J.N., SEIBLE, F. and CALVI, M.
Seismic design and retrofit of bridges, John Wiley &
Son, Inc., New York, 1996.
SEIBLE, F., PRIESTLEY, N., HEGEMMIER, G.A.
and INNAMORATO, D. Seismic retrofit of RC
columns with continuous carbon fiber jackets, Journal
of Composites for Construction, American Society of
Civil Engineers, Vol. 1, No. 2, May 1997, pp.52ff.
TRIANTAFILLOU, T.C. Seismic retrofitting of
structures with advanced composites A review,
Proceedings of FRP Composites in Civil Engineering,
Hong Kong, 2001, Volume 1, pp.807816.
SUTER. R., CONUS, F., PINZELLI, R. and CHANG,
K. Reinforcement of bridge piers with FRP sheets to
resist vehicle impact: tests on large concrete columns
reinforced with aramid sheets, Proceedings of FRP
Composites in Civil Engineering, Hong Kong, 2001,
Volume 1, pp.789796.
PATOARY, M.K.H. and TAN, K.H. Blast resistance of
prototype in-built masonry walls strengthened with
FRP systems, In: Tan, K.H. (Ed.), Proceedings of the
Sixth International Conference on FRP Reinforcement
for Concrete Structures, World Scientific, Singapore,
July 2003, Volume 2, pp.11891198.
CRAWFORD, J.E. and MORRILL, K.B. Retrofit
techniques using polymers and FRPs for preventing
injurious wall debris, In: Tan, K.H. (Ed.), Proceedings
of the Sixth International Conference on FRP Reinforcement for Concrete Structures, World Scientific,
Singapore, 2003, Volume 2, pp.11991208.
GALATI, N. Arching effect on masonry walls strengthened with FRP materials, MSc thesis, University of
Missouri-Rolla, USA, 2002, 124pp.

References

108. LI, T., SILVA, P.F., BELARBI, A. and NANNI, A.


Retrofit of unreinforced infill masonry walls with
FRP, In: Proceedings of the International Conference
Composites in Construction, Porto, Portugal, October
2001, pp.559563.
109. BLASCHKO, M., NIEDERMEIER, R. and ZILCH,
K. Bond failure modes of flexural members strengthened with FRP, In: Saadatmanesh, H. and Ehsani,
M.R. (Eds), Proceedings of the Second International
Conference on Composites in Infrastructure, University of Tucson, Arizona, 1998, Vol. 1, pp.315327.
110. TENG, J.G., CHEN, J.F., SMITH S.T. and LAM, L.
FRP Strengthened RC Structures, Chichester, Wiley,
2002.
111. NEUBAUER, U. and ROSTASY, F.S. Design aspects
of concrete structures strengthened with externally
bonded CFRP plates, Concrete and Composites, Proceedings of 7th International Conference on Structural
Faults and Repair, 1997, Vol. 2, pp.109118, ECS
Publications, Edinburgh.
112. DENTON, S.R., SHAVE, J.D. and PORTER, A.D.
Shear strengthening of reinforced concrete structures
using FRP composite, In: Hollaway, L.C., Chryssanthopoulos, M.K. and Moy, S.S.J. (Eds), Advanced
polymer composites for structural applications in construction, Woodhead Publishing Limited, 2004,
pp.134143.
113. IBELL, T.J., MORLEY, C.T. and MIDDLETON, C.R.
A plasticity approach to the assessment of shear in
concrete beam-and-slab bridges, The Structural Engineer, Vol. 75, No. 19, October 1997, pp.331338.
114. ESHWAR, N., IBELL, T.J. and NANNI, A. CFRP
strengthening of concrete bridges with curved soffits,
In: Forde, M.C. (Ed.), Proceedings of Tenth International Conference on Structural Faults and Repair,
Technics Press, Edinburgh, 2003, available as CD.
115. PORTER, A.D., DENTON, S.R. and RALPH, T.C.
The effect of curvature on strengthening with FRP, In:
Forde, M.C. (Ed.), Proceedings of Tenth International
Conference on Structural Faults and Repair, Engineering Technics Press, Edinburgh, 2003, available as
CD.
116. HORMANN, M., SEIBLE, F., KARBHARI, V. and
SEIM, W. Preliminary structural tests for strengthening of concrete slabs using FRP composites, Report
No. TR-98/13, University of California at San Diego,
1998.
117. HASSAN, T.K. and RIZKALLA, S.H. Bond
mechanism of NSM FRP bars for flexural strengthening of concrete structures, accepted for publication
ACI Structural Journal.
118. DE LORENZIS, L. and NANNI, A. Bond between
NSM FRP rods and concrete in structural strengthening, ACI Structural Journal, Vol. 99, No. 2,
MarchApril 2002, pp.123133.

119. DE LORENZIS, L., RIZZO, A. and NANNI, A. A


modified pull-out test for bond of NSM FRP rods in
concrete, Composites Part B: Engineering, Vol. 33,
No. 8, 2002, pp.589603.
120. DE LORENZIS, L. and NANNI, A. Characterization
of FRP rods as NSM reinforcement, Journal of Composites for Construction, Vol. 5, No. 2, May 2001,
pp.114121.
121. DE LORENZIS, L. and LA TEGOLA, A. Analytical
modelling of splitting bond failure for NSM FRP
reinforcement in concrete, In: Tan, K.H. (Ed.), Proceedings of the Sixth International Conference on FRP
Reinforcement for Concrete Structures, World
Scientific, Singapore, 2003, pp.975984.
122. MATTOCK, A.H. Rotational capacity of hinging
region in reinforced concrete beams, In: Proceedings
of International Symposium on Flexural Mechanics of
Reinforced Concrete, Miami, USA, 1964, pp.143181.
123. LEES, J.M. Flexure of concrete beams pre-tensioned
with aramid FRPs, PhD dissertation, University of
Cambridge, March 1997.
124. IBELL, T. and SILVA, P. A theoretical strategy for
moment redistribution in continuous FRP-strengthened concrete structures, In: Hollaway, L.C., Chryssanthopoulos, M.K. and Moy, S.S.J. (Eds), Advanced
polymer composites for structural applications in construction, Woodhead Publishing Limited, 2004,
pp.144151.
125. NAAMAN, A.E., PARK, S.Y., LOPEZ, M.M. and
TILL, R.D. Parameters influencing the flexural response of RC beams strengthened using CFRP sheets,
In: Burgoyne, C. (Ed.), Fibre-reinforced plastics for
reinforced concrete structures, Cambridge, UK, July
2001, Thomas Telford, Volume 1, pp.117125.
126. BURGOYNE, C.J. Rational use of advanced composites in concrete, In: Japan Concrete Institute,
Proceedings of the Third International Symposium on
Non-metallic FRP Reinforcement for Concrete Structures, Sapporo, Japan, 1997, Volume 1, pp.7588.
127. CASADEI, P., DENTON, S.R., IBELL, T.J. and
NANNI, A. Moment redistribution in continuous
CFRP-strengthened concrete members, In: Proceedings
of International Conference Composites in Construction 2003, Rende, Italy, September 2003.
128. EL-REFAIE, S.A., ASHOUR, A.F. and GARRITY,
S.W. Sagging and hogging strengthening of continuous reinforced concrete beams using CFRP sheets,
ACI Structural Journal, Vol. 100, No. 4, JulyAugust
2003, pp.446453.
129. DENTON, S.R. The strength of reinforced concrete
slabs and the implications of limited ductility, PhD
Thesis, Cambridge University, 2001.

85

Design guidance for strengthening concrete structures using fibre composite materials

130. SENTHILNATH, P., BELARBI, A. and MYERS, J.J.


Performance of CFRP strengthened reinforced concrete (RC) beams in the presence of delaminations and
lap splices under fatigue loading, In: Proceedings of
the International Conference on Composites in
Construction, Porto, Portugal, 1012 October 2001,
pp.323328.
131. MISSOURI DEPARTMENT OF TRANSPORTATION (MoDOT). Preservation of Missouris highway
infrastructure, Research Proposal Document, St
Louis, USA, 2003.
132. HEJLL, A., CAROLIN, A. and TLJSTEN, B.
Strengthening with CFRP under simulated live loads,
In: Tan, K.H. (Ed.), Proceedings of the Sixth International Conference on FRP Reinforcement for
Concrete Structures, World Scientific, Singapore,
2003, Volume 2, pp.13711380.
133. BARNES, R.A. and MAYS, G.C. The effect of traffic
vibration on adhesive curing during installation of
bonded external reinforcement, Proceedings of the
Institution of Civil Engineers Structures and Buildings, Vol. 146, No. 4, 2001, pp.403410.
134. CHEN, J.F. and TENG, J.G. A shear strength model for
FRP-strengthened RC beams, In: Burgoyne, C. (Ed.),
Fibre-reinforced plastics for reinforced concrete
structures, Cambridge, UK, July 2001, London,
Thomas Telford, Volume 1, pp.205214.
135. RICHART, F.E., BRANDTZAEG, A. and BROWN,
R.L. A study of the failure of concrete under combined
compressive stresses, Engineering Experimental
Station Bulletin No.185, 1928, University of Illinois.
136. HOWIE, I. and KARBHARI, V.M. Effect of tow sheet
composite wrap architecture on strengthening of
concrete due to confinement: I Experimental studies,
Journal of Reinforced Plastic and Composites, Vol.
14, No. 9, 1995, pp.10081030.
137. SAMAAN, M., MIRMIRAN, A. and SHAHAWY, M.
Model of concrete confined by fibre composites,
ASCE Journal of Structural Engineering, Vol. 24, No.
9, 1998, pp.10251031.
138. PICHER, F., ROCHETTE, P. and LABOSSIRE, P.
Confinement of cylinders and short concrete columns
with ACM, In: Saadatmanesh, H. and Ehsani, M.R.
(Eds), Proceedings of the First International Conference on Composite Infrastructures, University of
Tucson, Arizona, USA, January 1996, pp.829841.
139. SAAFI, M., TOUTANJI, H. and ZONGJIN, LI.
Behaviour of concrete columns confined with fibre
reinforced polymer tubes, ACI Materials Journal, Vol.
96, No. 4, 1999, pp.500509.
140. LILLISTONE, D. Non-ferrous compositely reinforced
concrete columns, PhD thesis submitted to the University of Southampton, 2000.

86

141. LAM, L. and TENG, J.G. Stressstrain model for FRPconfined concrete for design applications, In: Tan, K.H.
(Ed.), Proceedings of the Sixth International Conference on FRP Reinforcement for Concrete Structures,
World Scientific, Singapore, 2003, Volume 2,
pp.99110.
142. XIAO, Y. and WU, H. Compressive behaviour of
concrete confined by carbon fibre composite jackets,
ASCE Journal of Materials in Civil Engineering, Vol.
12, No. 2, 2000, pp.139146.
143. LILLISTONE, D. and JOLLY, C.K. An innovative
form of reinforcement for concrete columns using
advanced composites, The Structural Engineer, Vol.
78, No. 23/24, 5 December 2000.
144. CLARKE, J.L. and BIRJANDI, F.K. The behaviour of
reinforced concrete circular sections in shear, The
Structural Engineer, Vol. 7, No. 5, March 1993,
pp.7378 and 81.
145. FAM, A.Z., FLISAK, B. and RIZKALLA, S.H.
Experimental and analytical modelling of concretefilled fiber-reinforced polymer tubes subjected to
combined bending and axial loads, ACI Structural
Journal, Vol. 100, No. 4, 2003, pp.499509.
146. DENTON, S.R., CHRISTIE, T.J.C. and POWELL,
S.M. The development of design charts for column
strengthening using FRP, In: Ryall, M.J., Parke,
G.A.R. and Harding, J.E. (Eds), Proceedings of Fourth
International Conference Bridge Management: Inspection, Maintenance, Assessment and Repair, Thomas
Telford, 2000.
147. CUNINGHAME, J.R., JORDAN, R.W. and ASSEJEV,
A. Fibre reinforced plastic strengthening of bridge
supports to resist vehicle impact, Unpublished Project
Report, Transport Research Laboratory, Crowthorne,
1999.
148. LAM, L. and TENG, J.G. Design oriented stressstrain
model for FRP-confined concrete in rectangular
columns, Journal of Reinforced Plastics and Composites, Vol. 22, No. 13, 2003, pp.11491186.
149. MAALEJ, M. TANWONGSVAL, S. and PARAMASIVAM, P. Modelling of rectangular RC columns
strengthened with FRP, Cement and Concrete Composites, Vol. 25, 2003, pp.263276.
150. TENG, J.G. and LAM, L. Compressive behaviour of
carbon fibre reinforced polymer-confined concrete in
elliptical columns, ASCE Journal of Structural
Engineering, Vol. 128, No. 12, 2002, pp.15351543.
151. CAMPIONE, G., MIRAGLIA, N. and PAPIA, M.
Influence of section shape and wrapping technique on
the compressive behaviour of concrete columns confined with CFRP sheets, In: Proceedings of the International Conference on Composites in Construction,
Cosenza, Italy, September 2003, pp.301306.

References

152. TAN, K.W. Strength enhancement of rectangular


reinforced concrete columns using fibre-reinforced
polymer, Journal of Composites for Construction, Vol.
6, No. 3, 2002, pp.175183.
153. GARDEN, H.N. and HOLLAWAY, L.C. An experimental study of the failure modes of reinforced concrete beams strengthened with prestressed carbon
composite plates, Composites, Part B: Engineering,
Vol. 29B, 1998, pp.411424.
154. STOECKLIN, I. and MEIER, U. Strengthening of
concrete structures with prestressed and graduallyanchored CFRP strips, In: Tan, K.H. (Ed.), Proceedings of the Sixth International Conference on
FRP Reinforcement for Concrete Structures, World
Scientific, Singapore, 2003, Volume 2, pp.13211330.
155. EL-HACHA, R., GREEN, M. and WIGHT, G. Concrete beams post-strengthened with prestressed carbon
fibre reinforced polymer sheets, In: Proceedings of
International Conference Structural Faults and
Repair, London, 2001, Engineering Technics Press,
Edinburgh, 2001.
156. TJANDRA, R.A. and TAN, K.H. Strengthening of RC
beams with external FRP tendons, In: Tan, K.H. (Ed.),
Proceedings of the Sixth International Conference on
FRP Reinforcement for Concrete Structures, World
Scientific, Singapore, 2003, Volume 2, pp.985994.
157. KELLER, T. Strengthening of concrete bridges with
carbon cables and strips, In: Tan, K.H. (Ed.), Proceedings of the Sixth International Conference on FRP
Reinforcement for Concrete Structures, World Scientific, Singapore, 2003, Volume 2, pp.13311340.
158. BROWN, I.F. Abrasion and friction in parallel-lay
rope terminations, PhD thesis, University of Cambridge, UK, 1997.
159. ARAUJO, A.F. and GUIMARES, G.B. Moment
redistribution in continuous monolithic and segmental
concrete beams prestressed with external aramid
tendons, In: Tan, K.H. (Ed.), Proceedings of the Sixth
International Conference on FRP Reinforcement for
Concrete Structures, World Scientific, Singapore,
2003, Volume 2, pp.10031012.
160. DE LORENZIS, L., MILLER, B. and NANNI, A.
Bond of FRP laminates to concrete, ACI Materials
Journal, Vol. 98, No. 3, MayJune 2001, pp.256264.
161. HUTCHINSON, R.L. and RIZKALLA, S.H. Shear
strengthening of AASHTO bridge girders using CFRP
sheets, In: Proceedings of the Fourth International
Conference on FRP Reinforcement for Concrete
Structures, Baltimore, USA, 1999, pp.945958.
162. NURCHI, A., MATTHYS, S., TAERWE, L., SCARPA,
M. and JANSSENS, J. Tests on RC T-beams
strengthened in flexure with a glued and bolted CFRP
laminate, In: Tan, K.H. (Ed.), Proceedings of the Sixth
International Conference on FRP Reinforcement for
Concrete Structures, World Scientific, Singapore, July
2003, Volume 1, pp.297306.

163. TUMIALAN, J.G., GALATI, N., NAMBOORIMADATHIL, S.M. and NANNI, A. Strengthening of
masonry with FRP bars, In: Proceedings of Third
International Conference on Composites in Infrastructure, San Francisco, USA, June 2002, Available
as CD-ROM.
164. GERGELY, J. and YOUNG, D.T. Masonry wall
retrofitted with CFRP materials, In: Proceedings of
the International Conference Composites in Construction, Porto, Portugal, October 2001, pp.565569.
165. NORDIN, H. Flexural strengthening of concrete
structures with prestressed CFRP rods, PhD thesis,
Lule University of Technology, Sweden, 2003.
166. DE LORENZIS, L. and NANNI, A. Shear strengthening of reinforced concrete beams with near-surface
mounted fiber-reinforced polymer rods, ACI Structural
Journal, Vol. 98, No. 1, 2001, pp.6068.
167. CASADEI, P. and NANNI, A. SRP strengthening of
prestressed concrete double T-beams, RB2C report,
December, University of Missouri-Rolla, USA, 2003.
168. VALERIO, P. and IBELL, T.J. Shear strengthening of
concrete bridge decks using FRP bar, In: Tan, K.H.
(Ed.), Proceedings of the Sixth International
Conference on FRP Reinforcement for Concrete
Structures, World Scientific, Singapore, 2003, Volume
1, pp.539548.
169. KESSE, G. and LEES, J.M. Shear critical RC beams
strengthened with CFRP straps, In: Tan, K.H. (Ed.),
Proceedings of the Sixth International Conference on
FRP Reinforcement for Concrete Structures, World
Scientific, Singapore, 2003, Volume 1, pp.447456.
170. BANK, L.C. Report on Rapid Strengthening of RC
Bridges: Program SPR-0010 (36), Wisconsin Department of Transportation, USA, December 2002.
171. TLJSTEN, B. Strengthening of concrete structures
in torsion with FRP, In: Tan, K.H. (Ed.), Proceedings
of the Sixth International Conference on FRP Reinforcement for Concrete Structures, World Scientific,
Singapore, 2003, Volume 2, pp.11671176.
172. DESIDERIO, P. A maintenance strategy of FRP
strengthening systems, In: Tan, K.H. (Ed.), Proceedings of the Sixth International Conference on
FRP Reinforcement for Concrete Structures, World
Scientific, Singapore, 2003, Volume 2, pp.843852.
173. TOUTANJI, H. and DENG, Y. Flexural behaviour of
rc beams strengthened with carbon fibre sheets
bonded with organic and inorganic matrices, In:
Forde, M.C. (Ed.), Proceedings of Tenth International
Conference on Structural Faults and Repair 2003,
Engineering Technics Press, Edinburgh, 2003, available as CD.
174. Health and Safety at Work etc Act. London, HMSO,
1994.

87

Design guidance for strengthening concrete structures using fibre composite materials

175. BRITISH STANDARDS INSTITUTION. BS EN ISO


9002, Quality systems: Model for quality assurance in
production, installation and servicing, 1994, BSI,
London, 20pp.
176. BRITISH STANDARDS INSTITUTION. BS 1881:
Testing concrete, Part 207. Recommendations for the
assessment of concrete strength by near-to-surface
test, 1992, 20pp.

88

177. UNIVERSITY OF GLAMORGAN. Design of


bespoke testing equipment to evaluate lap shear test
on concrete specimens/FRP specimens, School of
Technology Report, November 1999.
178. BRITISH STANDARDS INSTITUTION. BS EN 923,
Adhesives terms and definitions, BSI, London, 1998.
179. AMERICAN SOCIETY FOR TESTING AND
MATERIALS. ASTM D 907. Standard terminology of
adhesives, American Society for Testing and Materials, Philadelphia, USA.

APPENDIX A

GLOSSARY OF TERMS
As many readers of this Report may be unfamiliar with fibre
composites and with adhesive technology, many of the terms
used are defined below. A more extensive glossary of adhesive terms is given in BS EN 923(178) and ASTM D 907(179).
Adhesive A polymeric material which is capable of holding
two materials together by surface attachment.
Aramid A manufactured fibre in which the fibre- forming
substance consists of a long-chain synthetic aromatic polyamide.
Bond The adhesion of one surface to another, with the use
of an adhesive or bonding agent.
Carbon fibre Fibres produced by the pyrolysis of organic
precursor fibres such as rayon, polyacrylonitrile (PAN) or
pitch in an inert atmosphere. The term is often used interchangeably with graphite. However, carbon fibres and graphite fibres differ in the temperature at which the fibres are
made and heat-treated, and the carbon content.
Composite or composite material A combination of highmodulus, high-strength and high-aspect-ratio fibre reinforcing material encapsulated by and acting in concert with a
polymeric matrix.
Cure To change the properties of an adhesive irreversibly
by chemical reaction into a more stable condition and to
develop the desired properties.
Epoxy resins Resins which may be of widely different
structures but which are characterised by the reaction of the
epoxy group to form a cross-linked hard resin.
Fabric Non-woven A textile structure produced by
bonding or interlocking of fibres, or both, accomplished by
mechanical, chemical, thermal or solvent means and combinations thereof.
Fabric Woven A generic material construction consisting of interlaced yarns or fibres, usually a planar structure.
Filament winding A reinforced plastics process that employs a series of continuous resin-impregnated fibres applied
to a mandrel in a predetermined geometrical relationship
under controlled tension.
Filler A relatively inert substance added to an adhesive to
alter its physical, mechanical, thermal, electrical or other
properties or to lower the cost.
FRP Fibre-Reinforced Plastics (or Polymers)
Glass fibre A fibre spun from an inorganic product of fusion
which has cooled to a rigid condition without crystallising.

Glass transition temperature (Tg) The approximate midpoint of the temperature range over which a polymeric adhesive changes from a relatively stiff and brittle material to a
viscous material.
Hand lay-up A process in which resin and reinforcement
are applied either to a mould or to a working surface and
successive layers built up by hand.
Hardener The curing agent or catalyst, which promotes
chemical cross-linking with the resin in two-component
adhesive systems.
Laminate A layer of fibre composite, either preformed or
formed in situ.
NSM Near-surface-mounted reinforcement.
Peel ply The outside layer of a reinforced plastic material,
which is removed to aid bonding.
Plate Preformed prismatic FRP plate, formed by pultrusion or manufacturing process, generally with all the fibres
arranged in the longitudinal direction.
Polymeric Adjective describing a material (most commonly organic) composed of molecules characterised by the
repetition of one or more types of monomeric units.
Pot life The period of time during which a multi-part adhesive can be used after mixing the components. (Note: The
pot life varies with the volume and temperature of the mixed
adhesive and the ambient temperature. The term pot life is
also used for the application of hot-melt adhesives for the
period for which an adhesive, ready for use, remains usable
when kept at normal operating temperature.)
Prepreg Reinforcing fibres in sheet or roll form impregnated with resin and stored for use.
Primer Material used to protect a surface prior to the
application of the adhesive, improve adhesion and/or
improve the durability or to stabilise/protect the substrate.
Pultrusion A continuous process for the manufacture of
composite profiles by pulling layers of fibres, impregnated
with a thermoset resin, through a heated die, thus forming
the ultimate shape of the profile.
Resin The reactive polymer base in adhesive and prepreg
matrix systems.
Substrate The material of the adherend adjacent to the
adhesive layer.
Thermoset A resin that is substantially infusible and
insoluble after being cured.
UHM Ultra high modulus.
Wet lay-up A method of making a reinforced product by
applying a liquid resin system while the reinforcement is put
in place, layer by layer.

89

APPENDIX B

SYSTEMS AVAILABLE IN THE UK


The following Tables give information on some of the
strengthening systems currently available in the UK from
supporters of the Project. The properties given are taken
from manufacturers and suppliers data sheets and are
thought to be correct at the time of publication. For design
purposes, actual properties must be obtained from the manu-

facturer. As test methods vary, the information should detail


the basis for the information (e.g. frequency of testing,
standard deviation).
In addition, strengthening materials are available from other
manufacturers and suppliers, who should be contacted for
the appropriate information.

Table B1: Suppliers of strengthening materials.

Supplier
Exchem

Trade name
Selfix Carbofibe

Degussa
(formerly MBT Feb)

MBrace
MBrace
MBrace
MBrace
Kevlar*
MBar
Tyfo Fibrwrap Systems

Fyfe

Sika

weber building solutions

Toray Europe Ltd.


* In association with Du Pont.

Sika CarboDur
Sika CarboDur DML
SikaWrap Hex 230C
SikaWrap Hex 100G
Enforce

Torayca UT70

Type of material
Carbon FRP plates
Carbon fibre rods
Carbon fibre sheet
Aramid fibre sheet
Glass fibre sheet
Carbon fibre sheet
Carbon FRP plates
Ultra high modulus carbon plates
Glass fibre sheet
Aramid fibre tape and sheet
Carbon fibre rods
Carbon fibre sheet
Carbon FRP plates
Carbon/aramid hybrid fibre sheet
Carbon bi-directional fibre sheet
Glass fibre sheet
Glass FRP plates
Glass/aramid hybrid fibre sheet
Glass bi-directional fibre sheet
Aramid fibre sheet
FRP anchors
Carbon FRP plates
UHM bespoke carbon fibre plates
Carbon fibre sheet
Glass fibre sheet
Carbon FRP plates
Carbon fibre sheet
Glass fibre sheet
Aramid fibre sheet
Ultra high modulus carbon FRP plates
Carbon fibre sheet

91

Design guidance for strengthening concrete structures using fibre composite materials

Table B2: Properties of fibre composite sheet materials.

Trade name
Enforce

Carbon
Carbon
Glass
Aramid

Strength
(N/mm2)
3800
2650
1700
2400

MBrace C130
MBrace C530
MBrace*
MBrace*
MBrace AR55
MBrace Kevlar
AK40, 60, 90
Selfix Carbofibe E

Carbon
Carbon
Carbon
Carbon
Glass
Aramid

3550
3000
3800
2650
1550
2100

235
390
240
640
74
120

Areal weight
(g/m2)
200
400
350
290, 420, 650,
850
300
300
200
400
915
320, 450, 650

Glass

3450

73

432

Effective
thickness (mm)
0.117
0.190
0.135
0.2, 0.29, 0.45,
0.59
0.11, 0.165
0.165
0.117
0.190
0.118
0.193, 0.286,
0.430
0.167

Selfix Carbofibe C

Carbon

4900

230

300

0.167

Selfix Carbofibe AR

Aramid

2900

100

240

0.167

SikaWrap 230C
SikaWrap 103C
SikaWrap 100G
SikaWrap 450A
SikaWrap 300A
Torayca UT70-20

Carbon
Carbon
Glass
Aramid
Aramid
Carbon

4100
3900
2300
2880
2880
4090

230
230
76
100
100
230

220
610
935
450
300
200

0.12
0.34
0.36
0.31
0.21
0.111

Torayca UT70-30

Carbon

4220

235

300

0.167

Tyfo SEH -51

Fibre

Modulus
(kN/mm2)
240
640
65
120

Width
(mm)
300
300
680
300
300, 500
300, 500
300
300
500, 1000
100, 300, 500
150, 300
& bespoke
150, 300
& bespoke
150, 300
& bespoke
300, 600
600
600
300
300
100, 250, 500,
1000
100, 250, 500,
1000
1372

Glass/Aramid
3238
72
915
0.36
hybrid
Tyfo SEH-51A
Glass
3238
72
915
0.36
1372
Tyfo SEH-25A
Glass
3238
72
505
0.19
1372
Tyfo WEB
Bidirectional
3238
72
295
0.116
1270
glass
Tyfo BC
Bidirectional
3238
72
813
0.32
1372
45 glass
Tyfo SCH-41
Carbon
3789
230
644
0.28
610
Tyfo SCH-41s
Carbon/Aramid
3789
230
644
0.28
610
glass
Tyfo SCH-11UP
Carbon
3789
230
298
0.127
610
Tyfo SCH-7UP
Carbon
3789
230
200
0.08
610
Tyfo CWEB
Bidirectional
3445
228
162
0.116
1270
carbon
Tyfo WAB
Aramid
3098
114
176
0.116
1270
Tyfo SAH-41
Aramid
3098
114
650
0.36
300
Notes: The properties are for dry fibres. The values should be treated as indicative only. See also note to Table B3.
*Available in Europe and the Middle East; not available in the UK.
The effective thickness is the total cross-sectional area of the fibres divided by the width of the sheet or tape.

92

Appendix B

Table B3: Properties of composite plate materials.

Trade name

Fibre

Enforce

Carbon

MBrace LM
MBrace MM
MBrace HM
MBrace UHM
MBrace 150*
MBrace 200*
Selfix Carbofibe S
Selfix Carbofibe M
Selfix Carbofibe H
Sika CarboDur S
Sika CarboDur M
Sika CarboDur DML
Tyfo UC
Tyfo UG

Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
Carbon
E-glass

Strength
(N/mm2)
28003000
24002600
Bespoke plates
1800
3100
2400
4500
28003000
24002600
2800
3200
1600
3100
3100
1110
2790
890

Modulus
(kN/mm2)
165
210
>300
140
170
250
150
165
210
150
200
280
165
210
360
230
41

Thickness
(mm)
1.2, 1.4
1.2, 1.4
Bespoke up to 50mm
1.2, 1.4
1.4
1.2, 1.4
Bespoke up to 48mm
1.2, 1.4
1.2, 1.4
1.2, 1.4 & bespoke
1.2, 1.4 & bespoke
1.2, 1.4 & bespoke
1.2, 1.4
1.4
Bespoke
1.2, 1.4, bespoke
1.2, 1.4, bespoke

Width
(mm)
10, 50, 80, 90, 100, 120
50, 80, 90, 100, 120
Bespoke up to 1m
50, 80, 100, 120
50, 80, 100, 120, 150
50, 80, 100, 120, 150
50, 100, 150, 200
10, 50, 80, 100, 120
10, 50, 80, 100, 120
50, 80, 120
50, 80, 120
50, 80, 120
50, 60, 80, 90, 100, 120, 150
50, 60, 90, 100
Bespoke
50, 100, bespoke
50, 100, bespoke

Note: These properties are taken from manufacturers data sheets and are thought to be correct at the time of publication
(2004). For design purposes, actual properties must be obtained from the manufacturer. As test methods vary, the information
should detail the basis for the information (e.g. frequency of testing, standard deviation).
*Available in Europe and the Middle East; not available in the UK.

Table B4: Properties of NSM rods and strips.

Supplier

Trade name

Degussa

Mbar Galileo
Mbar Leonardo
Aslan
Enforce

Hughes Brothers
weber building solutions

Dimensions
(mm)
7.5 dia.
7.5 dia.
2 x 16
1.4 x 10

Strength
N/mm2
2300
2000
2070
2800

Stiffness
kN/mm2
130
200
130
165

93

Design guidance for strengthening concrete structures using fibre composite materials

Table B5: Properties of epoxy adhesives.

Property

Supplier and trade name


Exchem
Fyfe

Degussa

Tensile strength (N/mm2)


Flexural strength
Shear strength

(N/mm2)

MBrace
laminate
adhesive

Resifix 31

32

24

23

51

47

19

>35

55

50

86

80

35

22

Currently
under test

6.5

Currently
under test

3.8

Currently
under test
Currently
under test

(N/mm2)

Flexural modulus (kN/mm2)

10

Shear modulus (kN/mm2)


Glass transition temperature, Tg, (C)

65

60

Selfix
Tyfo WS Tyfo TC
Carbofibe
Adhesive

weber building
Sika
solutions
Epoxy Plus SikaDur

30

18

15

2.2

2.1

9.8

12.8

93

82

60, 120

82

Notes: See note to Table B3.

Table B6: Properties of laminating resins.

Property

Tensile strength (N/mm2)


Flexural strength

weber
building
solutions

Supplier
Fyfe Tyfo S

Exchem

SikaDur
330

SikaDur
300

30

45

3.8
53

3
60

50

17

72

60

(N/mm2)

120

28

123

100

(kN/mm2)

3
55

5
60, 120

3
93

3
64

Flexural modulus
Glass transition temperature, Tg, (C)
Notes: See note to Table B3.

94

MBT

APPENDIX C

QUALITY CONTROL OF MATERIALS


This section is concerned with the quality control of the
materials used for strengthening. It is not intended to be a
specification, but covers some of the significant points that
should be included in a specification. The manufacturer
should supply characteristic values of the mechanical
properties to be used for design purposes (e.g. strength,
elastic modulus) which should be taken as the mean value
minus 2 standard deviations. Sufficient tests should be
carried out at regular intervals to ensure that this is
statistically valid.

Pultruded plates

Strengthening materials
General

All materials should be produced under an approved


quality scheme, such as ISO 9000.
All fibres, resins, composites and other materials should
be in accordance with the relevant ISO specifications,
Euronorms or other equivalent international standard.
Traceability of all materials should be ensured; all materials should be supplied with a certificate of conformity.
All external or independent testing should be carried out
in Approved Laboratories in accordance with international
standards or by the manufacturer under an approved
quality scheme.
All testing should be carried out in accordance with the
relevant ISO Standard, Euronorm or other internationally
accepted standard. Where no such standards exist, an
industry or company standard or method with a recognised history should be used.
Testing should consist of visual checks on the basic
materials (though this will provide limited information)
and, where appropriate, physical tests on the finished
elements as detailed below.

Fabric materials

The properties of a specified width of finished material


should be checked by testing samples.
The frequency of testing should be stated in the Quality
Plan. A minimum of one sample should be taken at the
start and finish of each production run.
The sample should be weighed to determine the weight
per square metre.
The elastic modulus and tensile strength should be
determined directly by testing the sample.

For pultrusion, the supply of fibres to the pultrusion line


should be monitored on a regular basis; this should be at
least once per hour.
All resins, hardeners, etc., should be used strictly in
accordance with the manufacturers instructions.
The speed of processing, processing temperature, etc.,
should be maintained within agreed limits and should be
checked and recorded regularly.
The properties of the plate should be checked by testing
samples.
The frequency of testing should be stated in the Quality
Plan. A minimum of one sample should be taken at the
start and finish of each production run.
The samples should also be checked for dimensional
accuracy.
Plates should be marked with a unique batch number at
regular intervals.
All pultruded plates should be supplied with a certificate
of conformity.

Prepreg plates

Samples of unidirectional fabric sheet may be made into


laminates, using the appropriate specified resin and the
laminate tested to determine the properties.
Where appropriate, properties should be determined in the
transverse direction as well as in the longitudinal direction.
All sheets and tapes should be supplied with a certificate
of conformity.
All individual rolls of material should be appropriately
labelled.

Cure schedules should be monitored and recorded.


Records should be maintained to demonstrate compliance
with the required procedure.
The finished product should be checked visually for
defects.
The finished product should be checked dimensionally
and by mass to ensure compliance.
Plates should be marked with a unique batch number.
All plates should be supplied with a certificate of
conformity.

Shells

Shells may be fabricated by filament winding or by hand


lay-up (or similar process) against a positive or negative
mould.

95

Design guidance for strengthening concrete structures using fibre composite materials

The finished product should be checked visually for


defects such as pin-holes and blisters.
The dimensions of the finished product should be checked.
The thickness should be measured at agreed locations.
Trial pieces should be made and tested. These may be
either sacrificial parts of the unit that can be removed for
testing or specially prepared samples made at the same
time and by the same process. In the latter case, the
samples may be formed on a flat surface to aid testing. In
both cases care must be taken to ensure that the sample is
representative of the material in the finished unit.
The frequency of testing should be stated in the Quality
Plan.
The samples should be weighed to check the density of
the product.
The samples should be tested in direct tension, in the
appropriate direction, to determine the elastic modulus
and the tensile strength.

All materials should be accompanied by a certificate of


conformity.
All materials, adhesives, laminating resins, etc., should be
stored and used strictly in accordance with the manufacturers instructions.
Accurate records should be maintained of all materials
used (e.g. delivery notes, batch numbers) and, where
required, the ambient conditions (e.g. temperature,
relative humidity).
Any independent testing that is required by the Client
should be carried out in Approved Laboratories in accordance with international standards or by the manufacturer
under his approved quality scheme.

Plates

Visual checks should be carried out on the plate to ensure


that the plate material is as specified.
The plate should be checked for local damage.
When bonded, the plate should be checked by tapping or
other means to ensure continuous adhesion.

Specials (shear straps, etc.)

Wet lay-up laminates

Specials may be fabricated by any appropriate method.


Testing should be in line with the recommendations for
shells.
Testing should be carried out on complete elements by an
appropriate method to determine the tensile strength and
elastic modulus.
The frequency of testing should be stated in the Quality
Plan.

Site requirements
As detailed in Chapter 10 and TR57, some site testing will
be required. The following points should be included in any
specification.

96

All work should be carried out in accordance with an


agreed Quality Plan, in accordance with ISO 9000 or
similar.

Visual checks should be carried out on mats, unidirectional tapes/fabrics, woven rovings and multi-axial fabrics
to ensure uniformity and conformity.
The completed laminate should be checked visually for
defects.
When required by the contract, trial pieces should be
made at the same time and by the same process. Care
should be taken to ensure that the trial pieces are representative of the material in the finished unit.
The frequency of testing should be as agreed with the
Client.
The samples should be tested to determine the elastic
modulus and the tensile strength.

APPENDIX D

SPECIALIST SUPPLIERS, CONTRACTORS, CONSULTANTS, UNIVERSITIES AND


OWNERS
This Appendix gives contact details of the organisations involved with the preparation of this Report. It is believed to be correct
at the time of going to press, but readers should check the information at the time of use.
Specialist suppliers
Degussa (formerly MBT Feb)
Albany House
Swinton Hall Road
Swinton
Manchester M27 4DT
Tel: +44(0)161 727 2731
Fax: +44(0)161 794 0944
www.mbtfeb.co.uk
Exchem Mining and Construction
Ventura Works
P O Box 7
Alfreton
Derbyshire DE55 7RE
Tel: +44(0)1773 540440
Fax: +44(0)1773 607638
e-mail: emc@exchem.com
www.exchem.com
Fyfe Company LLC
The Fibrwrap Company
Nancy Ridge Technology Center
6310 Nancy Ridge Drive Suite 103
San Diego
CA 92121-3209
USA
Tel: +1(1) 858 642 0694
Fax: +1(1) 858 642 0947
e-mail: info@fyfeco.com
www.fyfeco.com
Fyfe Europe Ltd
28 Aristippou Street
Glyfada 16674
Athens, Greece
Tel & Fax: +(30) 210 064 3402
e-mail: fyfe.europe@mail.gr

Fyfe Asia Pte. Ltd


10 Toh Guan Road #03-10 T.T.
International Tradepark
608838 Singapore
Tel: +65 95 898 5248
Fax: +65 95 898 5181
e-mail: fyfeasia@singnet.com.sg
Sika
Watchmead
Welwyn Garden City
Herts AL7 1BQ
Tel: +44(0)1707 394444
www.sika.co.uk
Toray Europe Ltd
3rd Floor
7 Old Park Lane
London W1Y 4AD
Tel: +44(0)20 7663 7779
Fax: +44(0)20 7663 7777
weber building solutions
Saint-Gobain Weber
Dickens House
Enterprise Way
Flitwick
Bedford MK45 5BY
Tel: +44(0)8703 330070
Fax: +44(0)1525 718988
e-mail:
mail@weberbuildingsolutions.co.uk
www.weberbuildingsolutions.co.uk

Specialist contractors
Concrete Repair Association
(Secretary Mr J Fairley)
Association House
99 West Street
Farnham
Surrey GU9 7EN
Tel: +44(0)1252 739145
Fax: +44(0)1252 739140
e-mail: info@associationhouse.org.uk
www.concreterepair.org.uk
Concrete Repairs Limited
Cathite House
23a Willow Lane
Mitcham
Surrey CR4 4TU
Tel: +44(0)20 8288 4848
Fax: +44(0)20 8288 4847
Email: mail@concrete-repairs.co.uk
www.concrete-repairs.co.uk
Specialist consultants
Arup Research & Development
13 Fitzroy Street
London W1T 4BQ
Tel: +44(0)20 7636 1531
Fax: +44(0)20 7755 3669
Email: RandD@arup.com
www.arup.com
Mouchel Parkman
West Hall
Parvis Road
West Byfleet KT14 6EX
Tel: +44(0)1932 337000
Fax: +44(0)1932 356122
www.mouchelparkman.com

97

Design guidance for strengthening concrete structures using fibre composite materials

Parsons Brinckerhoff Ltd


Queen Victoria House
Redland Hill
Bristol BS6 6US
Tel: +44(0)117 933 9300
e-mail: FRPcomposites@pbworld.com
www.pbworld.com
Tony Gee and Partners
TGP House
45-47 High Street
Cobham
Surrey KT11 3DP
Tel: +44(0)1932 868277
e-mail: ac@tgp.co.uk
www.tgp.co.uk

University of Bristol
Dept of Civil Engineering
Queens Building
University Walk
Bristol BS8 1TR
Tel: +44(0)117 928 7707
Fax: +44(0)177 928 7783
www.cen.bris.ac.uk
University of Glamorgan
School of Technology
Division of Built Environment
Pontypridd CF37 1DL
Tel: +44(0)1443 482121
Fax: +44(0)1443 482169
e.mail: dbtann@glam.ac.uk
www.glam.ac.uk/sot

Universities

Health & Safety Executive


Nuclear Safety Directorate
Room 304, Balliol Road
Bootle
Merseyside L20 3JZ
Highways Agency
Safety Standards & Research Division
Room 201, Heron House
49/53 Goldington Road
Bedford MK40 3LL
Tel: +44(0)1234 796107
Fax: +44(0)1234 796060
London Underground Ltd
84 Eccleston Square
London SW1V 1PX
Tel: +44(0)20 7027 9510

Owners
University of Bath
Department of Architecture & Civil
Engineering
Bath BA2 7AY
Tel: +44(0)1225 386908/6
Fax: +44(0)1255 386691
email: A.P.Darby@bath.ac.uk or
T.J.Ibell@bath.ac.uk
www.bath.ac.uk/ace

98

British Energy plc


Civil Design Group Engineering
Division
3 Redwood Crescent
Peel Park
East Kilbride G74 5PR
Tel: +44(0)1355 262000

Network Rail
Civil Engineering
40 Melton Street
London NW1 2EE
Tel: +44(0)20 7557 8000
Fax: +44(0)20 7557 9000
www.networkrail.co.uk
Oxfordshire County Council
Speedwell House
Speedwell Street
Oxford OX1 1NE
Tel: +44(0)1865 815 641

INDEX

adhesives 12-13, 69
curing 28, 41-2, 66, 73
definition 89
delamination 41
design considerations 28
design life 7
electrical currents and 10
flexural strengthening 31, 41-2
health and safety 12, 13
NSM reinforcement 37, 38, 73, 75, 77
prestressed FRP composites and 65, 66
properties 26, 28, 94
storage and site conditions 73
surface preparation 72, 73
temperature 10, 12, 28, 65, 73
testing 77, 80
thickness 72, 74, 75
workmanship and installation 12, 74-5,
76
alkali-resistant glass fibre 9, 27
anchorages
axial FRP 59
flexural strengthening 34, 36, 37, 67
material selection 13
maximum capacity 34, 47, 49
NSM reinforcement 37, 38-9, 41
prestressed FRP composites 65, 66-7
specials 12
tendons and U-wraps 67
wet lay-up laminates 67
see also bolted plate anchors; glass fibre
anchorages; mechanical fastening
aramid fibre, definition 89
aramid fibres and composites 9-10
application examples 17, 18, 20, 21, 22
blast protection 18
health and safety 13
impact loading 29
NSM reinforcement 21
partial safety factors 27
prestressing 65
ropes 67
shear strengthening 47
stress rupture 65, 66
suitability 14
assessment, structures 3-4, 24, 31-2, 71
axial FRP 49-50, 53, 57-9, 61
axial shortening 57
axially loaded members 53-64

beam-column connections 7, 12, 18, 58, 59


beams and slabs
application examples 16-17, 18-20
bridges 8, 18-20, 69
buildings 16-17
deep embedded bars 69
flexural strengthening 31-46
FRP advantages/disadvantages 5
NSM bars 68
shear strengthening 12, 17, 47-52, 67-8
wrapping 67, 73
see also moment capacity; shear capacity
bending moment capacity see moment
capacity
blast protection 14, 18, 29
bolted plate anchors 66, 67-8
see also mechanical fastening
bond, definition 89
bonding, electrical 10
bonds
columns 53
delamination 41
maximum anchorage capacity 47
non-destructive tests 77
NSM reinforcement 37, 73
prestressed FRP composites 65, 66
surface preparation 72-3
thick plates 39
see also adhesives; anchorages;
debonding; separation failure
bridges
application examples 7-8, 15-16, 18-21,
22, 28
appraisal 24
beams and slabs 8, 18-20, 69
deflection 40
deformation 40
design life 7
fatigue 41, 57
flexural strengthening 8, 41, 53
FRP advantages/disadvantages 5, 6
joints 21
partial safety factors 24
plates 8, 21
prestressed FRP composites 67
shear strengthening 7-8, 69
stress-strain model 55
wrapping 21
see also impact damage
brittle failure 25, 31, 56, 57, 68, 69
buildings 7, 15, 16-18, 40
see also beam-column connections;
beams and slabs; columns; structural
assessment; walls

car parks 1, 7, 17, 18


carbon fibre, definition 89
carbon fibres and composites 9-10, 14
application examples 7-8, 16-22, 28
costs 14, 18, 19
durability 26, 27
fabrics 18, 19, 20
health and safety 13
jackets 61
mats 19, 28
NSM reinforcement 11, 17, 19, 37, 68
partial safety factors 27
plates 11, 16-22, 66, 67
preformed shells 53
prestressed 65, 66, 69
rectangular columns 61
shear strengthening 17, 19, 20, 47, 68
sheets 16-22
stability 25
stress rupture 65, 66
strips 17, 18, 20
tapes 21, 69
chemical resistance, fibres 9
chimneys 21
see also towers
circular columns 17-18, 21, 29, 53-9, 62-4
classification, material properties 6
coastal structures see marine structures
coatings see over-coatings
collars 5, 59
columns 5, 53-64
application examples 17-18, 20-1, 22
preformed shells 11-12, 53
see also beam-column connections;
circular columns; moment capacity;
rectangular columns; shear capacity
compressive strengths 10, 53, 55-7
see also stress-strain model
concave surfaces, separation failures 34-5
concrete confinement 5, 17, 53-64
earthquake damage protection 14
preformed shells 11-12
torsional strengthening 69
see also wrapping
concrete preparation 72-3
concrete properties 25, 26
concrete splitting failure 37-8
concrete strength see concrete properties;
structural assessment
conformity see quality control
connections
preformed shells 12
see also beam-column connections; joints

99

Design guidance for strengthening concrete structures using fibre composite materials

consultants 97-8
contractors 71, 97
control samples see samples
corrosion 1, 4, 17, 21, 80
costs 4, 7-8
carbon systems 14, 18, 19
NSM reinforcement 1, 14, 37
preformed shells 12
cracks and cracking
application examples 18, 19, 21, 67
crack widths 25, 28, 40, 67
design considerations 28
flexural strengthening 40
inspection 80
NSM reinforcement 37, 38
prestressed carbon FRP straps 69
separation failure 34, 38
shear strengthening 49
curing
adhesives 28, 41-2, 66, 73
see also temperature
concrete repairs 71
definition 89

D
debonding
application example 17
bolted plate anchors and 67, 68
flexural strengthening 31, 39
inspection for 80
lap joints 56, 59
NSM reinforcement 37
shear strengthening 49, 56
surface preparation 72
wall strengthening 68
wrapped columns 59
deep embedded bars 69
deflection 17, 25, 28, 40-1
deformation 25, 27, 29, 40, 53, 57
delamination 41, 77, 80
design life 7
design modulus of elasticity 26-7
see also stiffness
design resistance moment, beams 32
design strain 27, 35
design strength 26, 27, 55
docks see marine structures
double-lap shear test 77
ductility 25, 40, 54
ductility enhancement 53, 54, 59, 61, 68
durability 6-7, 26, 27, 65
see also corrosion

E
earthquake damage
application examples 17, 18, 21, 22
protection against 14, 20, 28-9, 53
repair 18
economics see costs
electrical hazards 10, 14
electricity transmission poles 22
environmental aspects 6, 13, 73-4
see also health and safety
epoxy adhesives see adhesives

100

epoxy resins
definition 89
see also resins
extreme loadings 28-9

F
fabrics 10
adhesive application 75, 76
advantages 1
application examples 18, 19, 20-1
column strengthening 53
concrete preparation 72-3
installation 76
lap joints 56
properties 25
quality control 95
surface regularity and 35
see also multi-layer plates and fabrics;
woven fabrics
failure
design consideration 8
ductility enhancement and 59
FRP response 26
NSM reinforcement 37-8
post-tensioning systems 66
thick and multi-layer plates 39
see also brittle failure; deflection;
deformation; fatigue; fire protection;
lap joints; separation failure; shear
stress; splitting failure; stress rupture;
tensile rupture; vandalism
fan anchors 68
fatigue 25, 41, 57, 65
fibre composites
advantages/disadvantages 1, 5-7, 41
definition 89
health and safety 13
partial safety factors 27
post-tensioning 65
prestressed 65-7, 68, 69
properties 5, 9, 25, 92
shear strengthening 47
sheets 7, 8, 13, 27, 65, 92
strips 17, 18, 20, 49
tapes 21, 27, 69
tendons 65, 67
wet lay-up systems 13
see also aramid fibres and composites;
carbon fibres and composites; glass
fibres and composites; NSM (nearsurface-mounted) reinforcement
fibres 9-10, 13
filament winding 11, 27, 55, 89
fire damage 17
fire protection
adhesive selection 12
design considerations 10, 25, 26, 28
FRP advantages/disadvantages 6
over-coatings 77
flexural strengthening 31-46
anchorages 67
application examples 16, 17, 18, 20, 21
axial FRP 57-9, 61
beam-column connections 58

bridges 8, 41, 53
columns 53, 57-9, 61
materials 13
walls 68
see also NSM (near-surface-mounted)
reinforcement
full wrapping 47, 49, 69, 72

G
glass fibre, definition 89
glass fibre anchorages 68
glass fibres and composites 9-10
application examples 18, 20, 21
durability 26, 27
fabrics 20, 21
health and safety 13
NSM reinforcement 37, 68
partial safety factors 27
plates 21
preformed shells 18, 21, 53
prestressing 65
shear strengthening 47, 68
stress rupture 65, 66
suitability 14
see also filament winding
glass transition temperature (Tg) 12, 28, 89
glossary 89
grooves see slots
grouts 12, 21, 37, 68

H
hand lay-up 11, 27, 89
health and safety 10, 12, 13, 28, 71, 72
records 78, 79
see also environmental aspects; fire
protection
HM (high modulus) carbon systems 9, 14
hoop wrapping
application examples 20
columns 53, 55, 56-7, 58, 59, 61
earthquake damage protection 14
rectangular columns 29

I
identification 73, 78, 80
impact damage 6, 19, 26, 29
impact resistance 10, 20, 55
inorganic adhesives 69
inorganic resins 9
inspection 2, 7, 29, 71, 77, 79-80
see also maintenance; visual inspection
installation 5-6, 71-8
costs 7-8
NSM reinforcement 1, 37, 77
prestressed FRP composites and 65
surface regularity 34-5
see also quality control; workmanship
instrumentation 77, 79, 80

J
jetties see marine structures
joints
bridges 21
see also connections; lap joints

Index

laminating resins see resins


lap joints 5-6, 41, 56, 59, 76, 80
lateral deformation 53, 57
life-cycle costing see whole-life costing
life expectancy 69
lighthouses 21
limit state design 24-5
see also ultimate limit states
live loads, strengthening under 41-2

oval columns 61
over-coatings 10, 12, 25, 77-8, 80
overlaps see lap joints

M
maintenance 7, 8, 80
see also inspection
marine structures 1, 21-2
masonry see walls
materials 9-14
compatibility 80
design values 26-8
properties 6, 9, 25-6, 92-4
quality control 73, 95-6
storage 73
suppliers 91
testing 77
see also adhesives; concrete; fibre
composites; fibres; resins; SRP (steelreinforced polymer) materials
mechanical fastening 34, 66, 68, 69
moment capacity 3, 31-3, 58-9
see also beams and slabs; columns
moment redistribution 31, 40, 67
monitoring 2, 7, 29, 79-80
mortars see grouts
multi-layer plates and fabrics 39-40, 76

N
non-destructive tests 77
non-woven fabrics 10, 89
NSM (near-surface-mounted) reinforcement
1, 36-9
adhesives 37, 38, 73, 75, 77
anchorages 37, 38-9, 41
application examples 17, 19, 21
aramid fibres and composites 21
beams and slabs 68
carbon fibre and composites 11, 17, 19,
37, 68
costs 1, 14, 37
glass fibres and composites 37, 68
inspection 77
installation 1, 37, 77
post-tensioning 65
prestressed 68
shear strengthening 68
sizes and properties 11, 93
slots for 13, 38, 73, 77
suitability 13-14
vandalism and 29
wall strengthening 68
wet lay-up laminates 67
see also deep embedded bars
nuclear structures 7, 18, 28

P
paints see over-coatings
partial safety factors 26-7, 28
existing sections 24, 25, 32
stress-strain 58
tensile rupture 56
peel ply 19, 39, 74, 89
piers see marine structures
plates 5, 9, 11
application examples 8, 16-22
definition 89
fatigue 41
flexural strengthening 39-40
lap splices 41
partial safety factors 27
post-tensioning 65
prestressed FRP composites 66-7, 68
properties 11, 25, 38, 93
quality control 13, 95, 96
resins 11
stacking 39-40
suitability 1, 13
surface irregularities 35
tolerances 73
workmanship and installation 6, 13, 72-3,
74, 75-6
see also separation failure; specials;
warning plates
polyester/polyurethane adhesives 12
post-tensioning 65-6
preformed shells 11-12, 18, 21
columns 11-12, 53, 61
installation 71
partial safety factors 27
quality control 95-6
prepreg, definition 89
prepreg fabric 10, 21, 53
prepreg plates 11, 27, 95
prestressed FRP composites 65-7, 68, 69
prestressed structures 42
protective coatings see over-coatings
pull-off tests see testing
pultrusion 11, 89
see also plates

Q
quality control 13, 25, 71, 73, 95-6

R
railway structures 7, 10, 14, 19, 21, 22
see also tunnels
records 71, 73, 78, 79, 96
rectangular columns 53, 60-1
application examples 17, 18
impact loading 29
shear strengthening 47-52, 61
wrapping 29, 73

redistribution, moment 31, 40, 67


reinforcement 19, 25, 27-8, 73
see also corrosion; NSM (near-surfacemounted reinforcement); SRP (steelreinforced polymer) materials;
strengthening
repair
before strengthening 4, 40, 71
by FRP composites 17, 18, 19, 21
of FRP composites 80
whole-life costing 8
resin, definition 89
resins 9, 12-13
properties 26, 94
pultrusion process 11
sunlight and 10
workmanship and installation 74, 75
see also adhesives

S
safety see health and safety; partial safety
factors
samples 7, 25, 28, 77, 79, 80
see also testing
seismic loading see earthquake damage
separation failure 31, 33-6, 38, 45-6, 47
serviceability 40-2, 57
serviceability limit states 24-5, 26, 42, 65,
66
shape factor 60-1
shear capacity 31, 34, 47-8, 56, 65
see also beams and slabs; columns; shear
strengthening; shear stress
shear-crack-induced FRP separation 34, 38
shear strengthening 3, 5, 47-52
application examples 17, 18, 19, 20, 21
axial FRP 49-50, 57
beams and slabs 12, 17, 47-52, 67-8
bridges 7-8, 69
carbon fibre and composites 17, 19, 20,
47, 68
circular columns 53-9
debonding 49, 56
deep embedded bars 69
glass fibres and composites 47, 68
NSM reinforcement 68
prestressed carbon FRP straps 69
rectangular columns 47-52, 61
specials 12
ultimate limit states 48-9, 50, 56-7
walls 68
wet lay-up systems 13
wrapping 13, 47, 49, 56, 57, 67
shear stress 35-6, 47-8
bond failure 34
NSM reinforcement 38
prestressed FRP composites 66, 68
thick and multi-layer laminates 39
wet lay-up systems 13
see also shear capacity
shells see preformed shells
side-only wrapping 47, 49
site requirements 73-4, 96

101

Design guidance for strengthening concrete structures using fibre composite materials

slabs see beams and slabs


slots
useful perimeter 38
wet lay-up systems 67
workmanship 13, 37, 73, 75, 77
see also NSM (near-surface-mounted)
reinforcement
solar gain 78
spacing
plates 76
strips 49
specials 12, 96
splitting failure 37-8, 67
square columns see rectangular columns
SRP (steel-reinforced polymer) materials
68-9
stacking, plates 39-40
steel reinforcement see reinforcement
steel stress 27-8
stiffness
carbon systems 14, 26
design considerations 25, 26, 39, 55, 56
fibres 10, 26
plates 11, 38
preformed shells 12
woven fabrics 10
storage, materials 73
strain 26-7
flexural strengthening 31, 32, 33, 35
NSM separation failure 38
shear strengthening 48-9
tensile rupture 56
see also stress-strain model
straps 12, 69
see also strips
strengthening 1-8, 15-22
see also flexural strengthening; shear
strengthening; torsional strengthening
strengthening systems 91-4
stress rupture 12, 25, 41, 57, 65, 66
stress-strain model 26, 32, 53-5, 56, 57-8,
59
strips 17, 18, 20, 49
see also straps; tapes

102

structural assessment 3-4, 24, 31-2, 71


structural design 23-9, 71
substrate, definition 89
substrates 13, 31, 72-3, 74, 75
see also cracks and cracking; surface
irregularity-induced separation;
surface preparation; surface regularity
suppliers 97
surface irregularity-induced separation
34-5, 38
surface preparation 1, 12, 37, 72-3
surface regularity 34-5
symbols 23-4

T
tapes 21, 27, 69
see also straps; strips
temperature, adhesives 10, 12, 28, 65, 73
tendons 65, 67
tensile rupture 55-6
testing 7, 77, 79, 80
quality control 95
surface quality 73
wet lay-up systems 25
see also samples
thick plates 39-40
torsional strengthening 69
towers 8, 21, 22
tunnels 6, 12, 21, 28

U
U-wrapping 13, 47, 49, 67
UHM (ultra high modulus) carbon systems
9
ultimate limit states 24-5, 26, 27
flexural strengthening 31-2
NSM reinforcement 37
post-tensioning 66
separation failure 34, 35, 38
shear strengthening 48-9, 50, 56-7
stress-strain model 54, 56
tensile rupture 56
ultraviolet radiation 10, 12, 25, 77
underpasses 18

V
vandalism 6, 26, 29, 65, 77
vinyl ester adhesives 12
visual inspection 71, 74, 75-7, 79, 80, 96

W
walls 18, 29, 68
warning plates 29, 78, 80
water-cured glass fabric 21
wet lay-up, definition 89
wet lay-up systems
anchorages 67, 68
concrete durability 40
partial safety factors 27
quality control 96
suitability 13
testing 25
wrapping 13, 76
whole-life costing 4, 8
see also costs
workmanship 6, 12, 35, 40, 71-8
see also installation; quality control
worst credible strength 24, 25, 26
woven fabrics 10, 19, 26, 89
wrapping 3, 9
application examples 17, 18, 20-1, 22
beams and slabs 67, 73
bridges 21
impact loading 29
rectangular columns 29, 73
shear strengthening 13, 47, 49, 56, 57,
67
torsional strengthening 69
wet lay-up systems 13, 76
see also columns; concrete confinement;
fabrics; filament winding; full
wrapping; hoop wrapping; Uwrapping

You might also like